Untitled

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

MAY 2015 CHEN ET AL.

1849

Toward Improved Forecasts of Sea-Breeze Horizontal Convective Rolls at


Super High Resolutions. Part I: Configuration and Verification of a
Down-Scaling Simulation System (DS3)

GUIXING CHEN,* XINYUE ZHU, WEIMING SHA, AND TOSHIKI IWASAKI


Department of Geophysics, Graduate School of Science, Tohoku University, Sendai, Japan

HIROMU SEKO AND KAZUO SAITO


Meteorological Research Institute, Tsukuba, and Japan Agency for Marine-Earth Science and Technology,
Yokohama, Japan

HIRONORI IWAI AND SHOKEN ISHII


National Institute of Information and Communications Technology, Tokyo, Japan

(Manuscript received 23 June 2014, in final form 26 January 2015)

ABSTRACT

Horizontal convective rolls (HCRs) that develop in sea breezes greatly influence local weather in coastal
areas. In this study, the authors present a realistic simulation of sea-breeze HCRs over an urban-scale area at
a resolution of a few meters. An advanced Down-Scaling Simulation System (DS3) is built to derive the
analyzed data using a nonhydrostatic model and data assimilation scheme that drive a building-resolving
computational fluid dynamics (CFD) model. The mesoscale-analyzed data well capture the inland penetration
of the sea breeze in northeastern Japan. The CFD model reproduces the HCRs over Sendai Airport in terms
of their coastal initiation, inland growth, streamwise orientation, specific locations, roll wavelength, secondary
flows, and regional differences due to complex surfaces. The simulated HCRs agree fairly well with those
observed by dual-Doppler lidar and heliborne sensors. Both the simulation and observation analyses suggest
that roll updrafts typically originate in the narrow bands of low-speed streaks and warm air near the ground.
The HCRs are primarily driven and sustained by a combination of wind shear and buoyancy forces within the
slightly unstable sea-breeze layer. In contrast, the experiment without data assimilation exhibits a higher
deficiency in the reproduction of roll characteristics. The findings highlight that CFD modeling, given reliable
mesoscale weather and surface conditions, aids in high-precision forecasting of HCRs at unprecedented high
resolutions, which may help determine the roll structure, dynamics, and impacts on local weather.

1. Introduction (Kuo 1963; Asai 1970; LeMone 1973; Weckwerth et al.


1997). The rolls are aligned in long parallel lines that
Horizontal convective rolls (HCRs) are a common
highly influence the local distributions of wind and
feature of mesoscale shallow convection (Etling and
temperature. Roll convection also exhibits a coherent
Brown 1993; Atkinson and Zhang 1996; Young et al.
updraft/downdraft, which provides an efficient means
2002). HCRs typically form in the convective boundary
of vertically transporting momentum, heat, humidity,
layer (CBL) with moderately unstable stratification
and air pollutants in the boundary layer (e.g., LeMone
1976; Wyngaard and Moeng 1992; Weckwerth et al.
* Current affiliation: Department of Atmospheric Sciences, Sun
1996; Inagaki and Kanda 2010). A unique type of roll
Yat-sen University, Guangzhou, China. develops within the sea-breeze layer in coastal areas,
where cool marine air penetrates inland (Iwai et al.
2008; Oda et al. 2010). As sea breezes occur almost
Corresponding author address: Dr. Guixing Chen, Department
of Atmospheric Sciences, Sun Yat-sen University, 135 Xingangxi daily during summer and a majority of the population
Road, Guangzhou 510275, China. lives in coastal areas, further understanding and im-
E-mail: guixing_chen@yahoo.com proved modeling of sea-breeze HCRs are valuable for

DOI: 10.1175/MWR-D-14-00212.1

Ó 2015 American Meteorological Society


Unauthenticated | Downloaded 12/09/20 06:41 PM UTC
1850 MONTHLY WEATHER REVIEW VOLUME 143

forecasting local weather and for determining related et al. 2007; Maronga and Raasch 2013). To describe ap-
societal implications. propriate weather conditions, an option is to couple the
A general difficulty with numerical modeling of HCRs CFD model to a mesoscale model (Baik et al. 2009; Ashie
has been resolving their finescale structure and meso- and Kono 2011), potentially through advanced nesting
scale evolution. High-resolution calculations, such as strategies (Mirocha et al. 2014; Muñoz-Esparza et al.
large-eddy simulation (LES), are widely applied for 2014). In coastal areas, the variations in winds and tem-
explicit resolutions of roll convection at the scale of the perature can be significant because of the land–sea con-
CBL depth (Moeng 1986; Moeng and Sullivan 1994; trast (Miller et al. 2003). Therefore, to achieve realistic
Khanna and Brasseur 1998; Liu and Sang 2011). Sea- modeling of sea-breeze HCRs, coastal weather and sur-
breeze HCRs have a particularly small wavelength of face conditions should be precisely represented.
several hundred meters (Iwai et al. 2008; Oda et al. Considering the above-mentioned difficulties, fore-
2010). Thus, numerical models must have a grid spacing casts of sea-breeze HCRs are a challenge. In this study,
of tens of meters or smaller. At such a high resolution, we propose a solution with a numerical prediction sys-
the complicated surface conditions (buildings and steep tem based on advanced techniques. The system is built
topography) need to be described explicitly, for exam- to carry out two major tasks: represent the mesoscale
ple, by a computational fluid dynamics (CFD) model weather associated with roll formation and implement
(Kanda et al. 2004; Ashie and Kono 2011; Castillo et al. a building-resolving CFD simulation to describe in-
2011; Inagaki et al. 2012; Park and Baik 2013). More- dividual rolls over an urban-scale area. The goal is to
over, to elucidate the evolution of HCRs, numerical capture a specific sea-breeze event at a very high reso-
models not only need a spatial resolution fine enough to lution and to reproduce HCRs more realistically than
capture individual rolls, but also need to possess a domain those reported previously. In Part I of this study, we
size large enough to capture their mesoscale features evaluate the system performance by comparing nu-
(Atkinson and Zhang 1996). Such a high-resolution cal- merical results with intensive observations; in Part II
culation in a mesoscale domain requires a huge compu- (Chen et al. 2015, hereafter Part II), we examine the
tational resource, which has only become available in detailed impacts of land use and buildings on roll char-
recent years. These calculations allow for the simulation acteristics. Part I is organized as follows: Section 2 de-
of roll clouds during cold air outbreaks over the ocean scribes the configuration of the numerical prediction
(Liu et al. 2004; Gryschka and Raasch 2005; Eito et al. system, the experiment design, and the verification data.
2010) and of roll-like temperature patterns over urban Section 3 estimates the ability of the mesoscale-analyzed
areas (Ashie and Kono 2011). However, realistic urban- data to represent the inland progress of a sea breeze.
scale simulations of sea-breeze HCRs via building-resolving Section 4 extensively verifies the characteristics and
models are sparse. The roll structure, evolution, and structure of the CFD-simulated rolls. Section 5 presents
dynamics over coastal lands are not fully understood. a generation budget of the roll kinetic energy to verify
Another difficulty in simulating HCRs involves the the reproduction of roll-forcing mechanisms. Then, the
representation of roll formation environment. Roll char- conclusions are presented.
acteristics vary considerably for different weather condi-
tions and underlying surfaces. For instance, the roll aspect
2. Numerical system, experiment design, and
ratio (a ratio of the roll wavelength to the CBL depth)
observational data for verification
tends to increase with CBL instability, while the roll ori-
entation may change with CBL wind direction (e.g., Asai This section introduces the Down-Scaling Simulation
1970; Weckwerth et al. 1997). The structures of secondary System (DS3), which is designed for a high-precision
flows are also responsive to surface heating and the am- forecast of mesoscale weather at a super high resolution.
bient wind speed and direction (Raasch and Harbusch The system features one-way nesting with a parallelized
2001). As rolls evolve in a mesoscale area, they are likely CFD model coupled to a mesoscale model. In contrast to
subject to the impacts of changes in the atmospheric and other simple nesting systems, DS3 derives analyzed data
surface conditions. Many previous studies performed to describe the local weather using a convective-scale
simulations using idealized conditions or typical profiles data assimilation scheme in a mesoscale model, thereby
on uniform surfaces; thus, these studies were limited to potentially improving the accuracy of multiple down-
examining the overall statistics of roll behaviors (e.g., scaling processes. The system applies high-resolution
Moeng and Sullivan 1994; Khanna and Brasseur 1998). analysis data to obtain a short-range forecast of weather
Recent studies suggested that including inhomogeneous variables at even higher resolutions. Then, the system
surfaces in models is important for the realistic simulation employs a building-resolving CFD model to explicitly
of roll structures and their mesoscale evolution (Prabha describe complex surfaces and to simulate individual

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1851

HCRs over a suburban area near Sendai Airport. These TABLE 1. Configuration of the CFD model SIMPLERgo.
major components of DS3 are detailed in the following
Model parameters Specifications
subsections.
Basic equations Nonhydrostatic (compressible)
a. Meteorological model for mesoscale weather Coordinate Three-dimensional Cartesian
Discretization approach Finite-volume method
DS3 employs the Japan Meteorological Agency Grid system Staggered
Nonhydrostatic Model (JMA-NHM) to perform the Spatial resolution Adaptive/uniform
mesoscale calculations for the data assimilation and Time integration scheme Fully implicit
Advection scheme Third upwind scheme (QUICK)
extended forecast. The JMA-NHM is a community
Equation solver SIMPLER algorithm
model designed for operational forecasts and research Surface geometries treatment Blocking-off method
studies on mesoscale weather (Saito et al. 2006). The Turbulence scheme Lilly–Smagorinsky LES model
model is based on the fully compressible equations for
hybrid terrain-following coordinates. Many physical
processes, such as cloud microphysics, surface processes, Navier–Stokes equations are solved by an algorithm of
boundary layer scheme, convective parameterization, the Semi-Implicit Method for Pressure-Linked Equation
and atmospheric radiation, are included in the model. Revised code (SIMPLERgo; Patankar 1980; Sha et al.
The operational runs have provided numerical weather 1991). A full time-implicit scheme is applied; variable
predictions for Japan since September 2004 (Saito et al. fields are corrected iteratively until arriving at con-
2007). The model outperforms the preceding hydrostatic verged solutions at each time step. A high-order upwind
model in many aspects of mesoscale weather prediction. advection scheme Quadratic Upstream Interpolation for
For further details on the operational and research ap- Convective Kinematics (QUICK; Leonard 1979; Hayase
plications of the model see Saito (2012). et al. 1992) is employed. The discretization approach is
a finite-volume method on a staggered grid system. The
b. Data assimilation scheme
block-off technique is used to handle the complex ge-
The forecast accuracy of the mesoscale models is quite ometries such as buildings and topography (Patankar
sensitive to the initial conditions. One important method 1980; Sha 2002). For more detailed treatments of real land
for improving the quality of initial conditions is to link use and buildings see Part II. Regarding turbulence,
observations with the numerical models via data assimila- a classic Lilly–Smagorinsky LES is equipped to explicitly
tion. Miyoshi and Aranami (2006) established a data as- resolve large eddies and to parameterize small eddies
similation scheme for JMA-NHM using the local ensemble (Lilly 1962; Smagorinsky 1963). Moisture and radiation
transform Kalman filter (LETKF). Seko et al. (2011) processes are excluded. Using this model, Sha (2008)
modified the scheme to assimilate dense observations, such conducted a series of idealized experiments to illustrate
as radar radial winds and precipitable water vapor, to de- the impacts of steep orography, building arrays, and street
rive more accurate data. Seko et al. (2013) recently ex- blocks on local wind and temperature. The results from
tended the LETKF to a two-way nested system to obtain small-area calculations are encouraging for realistic ap-
analyzed data at the convective scale. The analyzed fields plications on an urban-scale domain in this study.
were used to improve the forecast of local intense rainfall.
d. Design of the numerical experiments for the
This nested system is incorporated into DS3 to obtain high-
sea-breeze event
resolution analyzed data. The system’s ability to represent
mesoscale weather in northeastern Japan is evaluated in The numerical experiments are set up to examine the
section 3 as one part of our verification effort. These sea-breeze event along the Pacific coast of northeastern
mesoscale-analyzed data are then used to initiate a short- Japan and the HCRs over Sendai Airport (Fig. 1).
range weather forecast at a higher resolution. Downscale simulations are run on five domains centered
at Sendai Airport (38.1358N, 140.948E), with horizontal
c. CFD model for local weather
resolutions of 15 km, 2 km, 400 m, 100 m, and 10 m, re-
For downscaling to a 10-m resolution, DS3 employs spectively. The calculations for domains 1–4 are im-
a local meteorological model based on a LES (Sha 2002). plemented using the JMA-NHM with topography from
The CFD model specifications are summarized in Table GTOPO30. For domain 5, the CFD modeling is per-
1 and are detailed as follows. The model is established in formed over realistic surfaces with high-resolution to-
three-dimensional Cartesian coordinates to resolve pography and GIS building information retrieved from
steep topography and buildings, rather than in terrain- the Zenrin Map Company.
following coordinates that are subject to severe trunca- The DS3 experiments with data assimilation (DS3_DA)
tion errors at large slope angles. The fully compressible are conducted over domains 1–2 using the nested

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1852 MONTHLY WEATHER REVIEW VOLUME 143

FIG. 1. Domain setup of the DS3 experiments. (a) Outer domain with a 15-km mesh, (b) inner domain with a 2-km
mesh, (c) downscale domain with a 400-m mesh, (d) downscale domain with a 100-m mesh, and (e) CFD model domain
with a 10-m mesh. In (a)–(d), the dashed rectangle denotes the next nesting domain. In (a),(b), the shaded area denotes
topographic altitudes. In (c),(d), the shaded area denotes land-use categories. In (e), the yellow blocks mark buildings.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1853

LETKF system (Figs. 1a,b). Both domains have 80 3 80 e. Intensive observations during the study period
horizontal grid points and 12 ensemble members. The
We focus on a typical sea-breeze event along the Pa-
member number is limited here considering the com-
cific coast of northeastern Japan. The event occurred in
promise between local computational resources and
the daytime on 19 June 2007, when thermally induced
real-time runs at a reasonable efficiency (Seko et al.
low pressure was established over the Japan Islands
2011, 2013). The forecast-analysis cycles of all members
under fair weather (Fig. 2a). By 1300 JST, the south-
start at 0300 Japan standard time (JST; JST 5 UTC 1
easterly sea breeze had penetrated most land surfaces
9 h) 18 June 2007 (;1.5 days ahead the validation time of
along Sendai Bay (Fig. 2b). The strong cooling of the sea
the HCRs). The operational pressure-level analysis data
breeze led to a low temperature of 238–258C, which was
(5-km mesh over Japan) within 3 days before the ex-
approximately 58C lower than that on the unaffected
periments are used for the initial conditions of 12
inlands. Within the surge of the cool marine air over the
members in domain 1. Surface parameters, such as al-
heated land, the HCRs formed during the afternoon
bedo and wetness, are set using 100-m mesh land-use
hours (Iwai et al. 2008; Oda et al. 2010).
data from the Geospatial Information Authority of Ja-
In June 2007, an observation campaign on sea breezes
pan. The unperturbed boundary condition is applied to
was conducted over Sendai Airport. In addition to
domain 1, and a covariance inflation of 1.2 is used in the
conventional surface observations, intensive observa-
LETKF system. Data assimilation is conducted in a 6-h
tions from ground-based NICT and Electronic Navigation
cycle using conventional hourly observations, such as
Research Institute (ENRI) Doppler lidars (Komatsubara
sounding profiles and surface pressure. The ensemble
and Kaku 2005) and Japan Aerospace Exploration
forecasts of domain 1 provide initial/boundary condi-
Agency (JAXA) helicopter flights (Matayoshi et al. 2005)
tions for domain 2. The data assimilation in domain 2 is
were obtained. The lidar locations, helicopter flights,
conducted in 1-h cycle using 10-min surface records of
and measuring equipment were detailed in an early
pressure and winds in northeastern Japan, as well as
report by Iwai et al. (2008). The NICT and ENRI lidars
30-min wind profiles derived from the velocity–azimuth
have range resolutions of 90 and 30 m, respectively.
display (VAD) analysis of the National Institute of In-
Using these intensive observation data, we attempt to
formation and Communications Technology (NICT)
evaluate the mesoscale-analyzed data’s representation
lidar over Sendai Airport (Ishii et al. 2007). To highlight
of the sea-breeze intrusion during the daytime on
the effect of the analyzed data in DS3_DA, we conduct
19 June. For the observed surface conditions, we use
a parallel run without data assimilation (DS3_noDA).
10-min records of winds and temperature from the JMA
All parameters in DS3_noDA are the same as those in
Automated Meteorological Data Acquisition System
DS3_DA described above, except that no observational
(AMeDAS). For the vertical structure of the sea breeze,
data are input to the LETKF system in domains 1–2.
we use the 30-min wind profile from the VAD analysis of
Note that DS3_noDA can somewhat present the
the NICT lidar. We also verify the CFD-simulated
weather conditions but not as accurately as DS3_DA
HCRs over Sendai Airport at 1305 JST. To observe
because the initial and boundary conditions of domain 1
the structure of the sea-breeze HCRs, we use the radial
are given by operational analysis data.
velocities of 1-min plane position indicator (PPI) scans
The latest analysis data from domain 2 can be used to
from the ENRI lidar and the retrieved three-
initiate an extended short-range forecast in domains 3–4
dimensional winds from the dual-Doppler lidar. These
(Figs. 1c,d). Domain 3 (4) has 80 3 80 (120 3 120)
observational data are compared with the CFD simu-
horizontal grids and a forecast time of 1 h. The output
lations to verify the roll characteristics and structure. We
from domain 4 offers initial and boundary conditions for
also use high-resolution (;1.16 m) winds and tempera-
the CFD modeling in domain 5 for a horizontal size of
ture from helicopter measurements to verify the de-
10 km and a vertical extent of 500 m at a uniform 10-m
tailed dynamic and thermodynamic properties of the
mesh (Fig. 1e). In the CFD model, both the lateral and
HCRs, as well as the associated roll-forcing mechanisms.
upper boundary conditions are handled by an improved
Orlanski radiation condition that is free of reflection
from gravity wave propagation (Orlanski 1976), while 3. Verification of the sea-breeze event over
the bottom boundary condition (ground temperature) is northeastern Japan
fixed because it minimally changes over a short period.
a. Inland penetration of the sea breeze
The building wall temperature is set to the prescribed
ground temperature from domain 4. The time step is 1 s. We compare the hourly data from the DS3 experi-
The forecast time is 600 s, which is approximately 4–5 ments with the AMeDAS records to verify the system’s
large-eddy turnover times. representation of mesoscale weather. Figure 3 shows the

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1854 MONTHLY WEATHER REVIEW VOLUME 143

FIG. 2. A sea-breeze event over northeastern Japan on 19 Jun 2007. (a) Analysis
chart of the mean sea level pressure at 1200 JST; and (b) the surface temperature and
winds at 1300 JST as observed by AMeDAS. A full (half) wind barb is 2 (1) m s21. In
(b), the topography is shaded in gray.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1855

FIG. 3. Hourly variations in the surface wind and temperature trend. (a) AMeDAS observation, (b) analyzed data in
DS3_DA, and (c) forecast in DS3_noDA. The temperature trend is estimated as a temperature difference between the valid
hour and the following hour. In (a), the sea-breeze front at each valid hour is marked by the dashed line based on its passage
over the AMeDAS sites. In (b) and (c), a contour of 0.28C h21 outlines the approximate sea-breeze front, where the cooling
due to sea breeze is comparable to the surface heating due to solar radiation. A full (half) wind barb is 2 (1) m s21.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1856 MONTHLY WEATHER REVIEW VOLUME 143

progress of the sea breeze on 19 June 2007. At 0900 JST, experienced over most of the lowlands on the lee side
the onset of the sea breeze is recorded at Sendai Airport, (east) of the Ou Mountains. Both DS3 experiments
where a southeasterly wind of 4 m s21 is followed by produce a cooling pattern (Figs. 3b,c) very similar to the
a temperature drop of 20.58C h21 (Fig. 3a). The tem- surface observations (Fig. 3a), except that DS3_noDA
perature at the other AMeDAS sites increases by ap- simulates a weaker cooling rate to the west of Sendai
proximately 18C h21. The numerical experiments of Airport. From the overview in Fig. 3, it seems clear that
both DS3_DA and DS3_noDA capture a narrow zone of the downscale experiment (DS3_noDA) can offer
cooling along the coastline and extensive warming over a reasonably good forecast, while the experiment with
land (Figs. 3b,c). At 1000 JST, the sea breeze forms over data assimilation (DS3_DA) better captures the sea-
most of the coastal areas and begins to penetrate inland breeze event in terms of its timing and cooling rate.
(Fig. 3a). The progress is relatively fast to the west of the
b. Atmospheric conditions over Sendai Airport
airport, with an inland extent of ;16 km from the
coastline. A cooling rate of up to 218C h21 occurs along The local weather over Sendai Airport is further
the coastal areas. At this time, DS3_DA captures the evaluated by comparing AMeDAS records with ana-
sea-breeze progress in terms of its penetration distance lyzed data. Figure 4 shows that the observed tempera-
and cooling rate near the airport (Fig. 3b). In contrast, ture increases rapidly after sunrise and reaches a peak of
DS3_noDA presents a relatively slow progress of the sea 23.28C at 0900 JST. Even under fair weather, the tem-
breeze and a weak cooling rate (Fig. 3c). perature decreases in the daytime because of the sea
At 1100–1200 JST, the sea breeze continues to pene- breeze, except for a small rebound at 1200 JST. Both
trate inland perpendicular to the coastline (Fig. 3a). The DS3 experiments reproduce the rapid temperature in-
progress is relatively fast to the northwest of Sendai crease after sunrise at 0600 JST but at a smaller ampli-
Airport, and the sea breeze gradually arrives at the tude than the observations. DS3_DA produces both the
foothills of the Ou Mountains. An enhanced easterly or temperature maximum at 0900 JST and the secondary
southeasterly wind follows the passage of the sea-breeze minimum at 1100 JST, which are consistent with the
front, with the strongest cooling of approximately observations. DS3_noDA, however, produces a delayed
228C h21 at the leading edge. At coastal areas near temperature peak at 1100 JST. Note that the wind speed
Sendai Airport, the sea breeze increases to 4–6 m s21 from DS3_DA is ;1 m s21 stronger than that from
and induces a gentle but continuous cooling. Such fea- DS3_noDA and becomes closer to the observations. The
tures are well captured by the analyzed data in DS3_DA well-simulated temperature and winds indicate that
(Fig. 3b). In contrast, DS3_noDA presents a delayed DS3_DA produces a superior representation of the sea
cooling at the leading edge of the sea-breeze front breeze strength.
(Fig. 3c). Over coastal areas to the south of the airport, For accessing the vertical structure of sea breeze, the
both numerical experiments simulate a sea-breeze prog- profiles of horizontal winds by the DS3 experiments are
ress that is slower than the observed progress, probably compared with those obtained using the 208 elevation
due to the orographic effects of coastal mountains. VAD scan from the NICT lidar over Sendai Airport.
For 3 h (1000–1200 JST), the sea breeze penetrates Figure 5a shows that in the early morning before
30–35 km inland in both the AMeDAS observations and 0700 JST, the observed wind vectors are northerly above
DS3 experiments; this movement indicates a speed of 1000 m AGL (above ground level) and northwesterly in
10–12 km h21. As suggested by Simpson (1969) and Sha the lower layer. In the morning after 1000 JST, the
et al. (1991), the sea breeze is a typical gravity current, southeasterly wind becomes dominant in the layer be-
whose speed can be given by U ’ 0.62(gDDT/T)1/2, low 500 m AGL, while the westerly wind prevails above
where g is the gravitational acceleration, D is the depth 1000 m AGL as the return flow of the sea-breeze circu-
of cool marine air, T is the temperature of cool marine lation. In the afternoon, the depth of the low-level
air, and DT is the temperature difference between ma- southeasterly wind continues to increase over time,
rine and inland air. Given D 5 220 m from the lidar while the maximum wind speed remains at ;200 m
analysis (Iwai et al. 2008; Oda et al. 2010) and DT 5 3– AGL and the wind direction below only slightly
4 K from the AMeDAS records at 1000–1200 JST, the changes, as also indicated by Iwai et al. (2008). Such
theoretical U is estimated as 10–12 km h21. Therefore, a wind pattern, with speed shear and little directional
the observed and simulated progress speeds in Fig. 3 are shear in the CBL, is favorable for roll formation and
consistent with the theoretical analysis. sustenance (e.g., Weckwerth et al. 1997).
At 1300 JST, the mature phase of the sea breeze is Figure 5b shows that DS3_DA well reproduces the
characterized by a strong surge of southeasterly flow establishment of the sea-breeze circulation in the day-
from the Pacific coast (Fig. 3a). Extensive cooling is time. The analyzed data also display a gradual

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1857

FIG. 4. Hourly variations in the near-surface wind at 10 m AGL and the temperature at 2 m
AGL over Sendai Airport on 19 Jun 2007. A full (half) wind barb is 2 (1) m s21.

deepening of low-level southeasterly wind in the after- initiate early in the coastal villages (A6 and A7). In
noon, consistent with the lidar observations. Figure 5c contrast, the rolls are relatively suppressed or absent
shows that DS3_noDA presents a persistent northwest- from the rice paddy fields (A3 and A5).
erly wind above 1000 m AGL. The run reproduces Figure 6b shows that the simulated HCRs, based on
a southeasterly sea breeze near the surface at 0900– the forecast by DS3_noDA, also form at coastal areas
1100 JST but with a weaker wind speed than that of the and develop over land. The rolls are aligned from
observations and DS3_DA. These differences between southeast to northwest but with a more northward
DS3_DA and DS3_noDA suggest that data assimilation component than those in Fig. 6a. The specific roll lo-
improves the representation of temperature and wind cations over the airport and inland regions in Fig. 6b
variations associated with the sea-breeze event over seem to differ from those in Fig. 6a. For instance, the
Sendai Airport. rolls over area A4 are shifted northward and are lo-
cated near the rolls over area A2. The rolls over areas
A2/A4 and downstream are more widely spaced than
4. Verification of the sea-breeze HCRs over those in Fig. 6a. This finding is somewhat expected
Sendai Airport because different initial/boundary conditions are ap-
plied in the two CFD experiments. To estimate fore-
a. General features of the CFD-simulated HCRs
cast accuracy, the simulated rolls are compared with
For an overview of the roll characteristics, we examine the observations. As indicated, we focus our verifica-
the spatial pattern and temporal evolution of the CFD- tion on the roll orientation, location, and wavelength,
simulated HCRs. Figure 6a shows a full-domain snap- and the associated regional differences over Sendai
shot of the simulated vertical motion by DS3_DA. It is Airport.
clear that vertical motion exhibits a streaky structure; Figure 7a shows the temporal evolution of the vertical
parallel convective bands are oriented from southeast to velocity at X 5 3 km in DS3_DA. At the initial time step,
northwest and are spaced several hundreds of meters the condition corresponds to a forecast of the mesoscale
apart. All of the HCRs form over land at a distance of model at 100-m mesh. The updrafts have a magnitude
1 km or more from the coastline. The major rolls con- within 60.6 m s21 and behave like rolls with nearly
tinue to grow downstream and stretch several kilome- regular spacing. The updrafts are much weaker than the
ters inland. The rolls are more evident downstream of observed updrafts as shown later; thus, the rolls are
the major buildup areas (A1, A2, and A4); some of them partly resolved in the mesoscale model. A possible cause

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1858 MONTHLY WEATHER REVIEW VOLUME 143

FIG. 6. Vertical velocity at 100 m AGL over a full domain of the


CFD model at a forecast time of 300 s (valid at 1305 JST 19 Jun
2007). (a) Simulation based on 1-h forecast from the analyzed data
of DS3_DA. (b) As in (a), but based on the forecast from
DS3_noDA. Buildings are plotted in the gray contour. The major
distributions of land use in the vicinity of Sendai Airport are labeled
as follows: airport terminal (A1), industry compounds (A2), rice
paddy fields (A3 and A5), residential village (A4), and coastal
villages (A6 and A7).

FIG. 5. Hourly variations in the vertical profile of the horizontal


winds over Sendai Airport during 0400–1500 JST 19 Jun 2007.
is that mesoscale model, with a resolution ranging from
(a) VAD analysis of the NICT lidar from 208 elevation scans, ;O(1) km to ;O(100) m, may fall within the ‘‘gray
(b) analyzed data in DS3_DA, and (c) forecast in DS3_noDA. zone’’ or ‘‘terra incognita,’’ in which neither the 1D
vertical diffusion in boundary layer scheme nor the
subgrid-scale LES closure are appropriate (Wyngaard
2004; Dorrestijn et al. 2013; Shin and Hong 2013; Beare

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1859

FIG. 7. Temporal variations in (a) the vertical velocity at X 5 3 km and (b) the turbulent
energy spectrum averaged over an area of X 5 3–5 km and Y 5 0.5–9.5 km. The estimate is
made for a south–north cross section at 100 m AGL from the DS3_DA experiment as shown in
Fig. 6a. In (b), the wavenumber has been multiplied by the Y distance along which the spectrum
is estimated. The dashed line denotes the K25/3 inertial range slope.

2014). The following series in the CFD model show that rolls become well developed after a forecast time of
the updrafts become strong and exhibit regional fea- ;300 s, and the CFD simulation reaches a quasi–steady
tures. From the initiation of weak and uniform pertur- state (Fig. 7b). The time scale is substantially shorter
bations, the CFD model gradually reproduces the than that in other studies on a typical CBL (1–2 km
strengthened updrafts, particularly downstream of the deep), because the sea breeze layer is very shallow
village and airport runway (Y 5 3.8 and 5.5–8 km). Some (;200 m) in this particular case. Similar growth of the
updrafts are displaced from their initial positions, re- rolls and turbulent energy occur in DS3_noDA (figure
sulting in an uneven spacing between the rolls. An omitted). In sections 4b–d, the instantaneous fields at
analysis of the turbulence spectra also suggests that the that time step are discussed.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1860 MONTHLY WEATHER REVIEW VOLUME 143

b. Roll orientation, location, wavelength, and DS3_DA produces two rolls shifted slightly to the south
regional differences (Figs. 8c,d), while that from DS3_noDA yields two rolls
shifted northward with respect to the observations
Figure 8 shows the horizontal analysis of the wind (Figs. 8e,f). The shifted locations correspond well to the
fields over Sendai Airport. The lidar-observed HCRs deviated wind direction and roll orientation in the two
are identified as the bands (CC0 –JJ0 ) of coherent up- CFD simulations. These rolls can be traced back to
drafts (Figs. 8a,b). The bands extend from southeast to coastal village A7 (Figs. 6a,b). During its inland evolu-
northwest, parallel to the mean wind direction. The tion, the southern roll mainly passes village A4, while
CFD-simulated HCRs in the two CFD experiments are the northern roll extends over rice paddy field A3. Both
also aligned streamwise (Figs. 8c–f). In particular, the simulations reproduce a relatively strong updraft for the
rolls from DS3_DA exhibit an orientation almost iden- southern roll that corresponds to the observed roll CC0
tical to the observed orientation, while those from (Figs. 8c–f). The warm surfaces of buildup areas may be
DS3_noDA are oriented more northward. The differ- conducive for roll growth and sustenance (Raasch and
ence in roll orientations corresponds well to the average Harbusch 2001; Miao and Chen 2008; Ashie and Kono
wind direction, which has an azimuthal angle of ;1358 2011). The underlying physics are further investigated in
(lidar observations), ;1338 (DS3_DA), and ;1398 Part II.
(DS3_noDA). Figure 8 also shows that the roll orienta- In the northern zone of the dual-Doppler lidar anal-
tion change with height is small from 50 to 100 m AGL in ysis domain, there are four rolls GG0 –JJ0 observed in
both the observations and simulations. Such a wind- Figs. 8a and 8b. The simulation from DS3_DA well re-
parallel roll orientation is expected when the CBL produces roll GG0 . Although the simulated rolls HH0 –II0
directional shear is small and when the speed shear is match the observations at the upstream locations, they
directed downwind (e.g., Asai 1970; LeMone 1973; tend to shift westward on the downstream side, likely
Weckwerth et al. 1997). A comparison with the next because the mean flow is more westward than observed.
volume of PPI scans 12 min later (figure omitted) in- Roll JJ0 is weak in the observations and is barely dis-
dicates that these observed rolls are persistent and tinguishable in DS3_DA. DS3_noDA also has a de-
nearly steady, except for a slight displacement toward ficiency in simulating the locations of rolls GG0 –JJ0
the northeast (Iwai et al. 2008) likely due to a slow (Figs. 8e,f). The simulated rolls generally have a north-
veering wind direction (Figs. 4–5). ward shift following the mean wind direction. On the
In addition to the roll orientation, the lidar measure- other hand, these rolls observed in the northern zone are
ments allow us to verify the specific locations of in- somewhat weaker than those in the southern zone
dividual rolls. In the southern zone of the dual-Doppler (Fig. 8b). Such a gradient of vertical velocity is not
lidar analysis domain, there are two elongated rolls ob- simulated immediately at the corresponding zones in the
served near the lidar sites, as marked by EE0 and FF0 CFD model, but it becomes clear in the downstream
(Figs. 8a,b). The roll locations are reproduced well by areas near the NICT site (Figs. 8d,f). The simulated rolls
the CFD simulation from DS3_DA (Figs. 8c,d). The that are well defined near the NICT site appear to be
CFD simulation from DS3_noDA also captures the lo- stronger than those detected by lidar, likely because of
cation of roll FF0 (Figs. 8e,f). Although the roll EE0 is the effectively lower resolution of the lidar retrieval.
reproduced on the upstream side, it moves northward Figure 8 also shows that the rolls are spaced at regular
and merges with its northern neighbor. As a result, the intervals, which enables us to estimate the extent to
roll location does not match the observed roll EE0 on the which the CFD simulations reproduce the roll wave-
downstream side. A full-domain view of Figs. 6a and 6b length. In the northern zone of the dual-Doppler anal-
shows that the rolls EE0 and FF0 originate from the ysis, the observed rolls display a mean spacing of ;450 m
coastal buildup area A6. The corresponding newly (Figs. 8a,b). In the southern zone, the mean spacing
formed rolls share similar locations in the two CFD between two major rolls (EE0 and FF0 ) is ;515 m. The
simulations, possibly because they are initially disturbed wavelength and regional difference are consistent with
by the same underlying surfaces. As the rolls evolve those estimated by power spectral analysis (Iwai et al.
inland with slightly different orientations in the two 2008). Figures 8c and 8d show that the simulated rolls
simulations, their specific locations gradually differ in from DS3_DA have a wavelength of ;415 m in the
the Sendai Airport area (Figs. 8c–f). northern zone and of ;475 m in the southern zone; these
In the southernmost corner of the dual-Doppler lidar values are slightly smaller than the observed values. The
analysis domain, there is a signature of two other rolls difference in roll wavelength between the two zones is
(CC0 and DD0 ), with stronger updrafts observed in the ;60 m, which is nearly the same as the lidar observa-
southern roll CC0 (Figs. 8a,b). The CFD modeling from tions. Figures 8e and 8f show that, in the simulation from

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1861

FIG. 8. Vertical velocity and horizontal winds over Sendai Airport at 1305 JST 19 Jun 2007. (a),(b) Dual-Doppler
lidar retrieval; (c),(d) CFD simulation from DS3_DA; and (e),(f) CFD simulation from DS3_noDA. The lines
CC0 –JJ0 mark the axis of the observed roll updrafts. In (b), the dashed line marks the helicopter flight in Fig. 11. In
(c), the dashed line AB marks the cross-roll section in Fig. 10.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1862 MONTHLY WEATHER REVIEW VOLUME 143

DS3_noDA, the rolls have a wavelength of ;400 m in


the northern zone. However, the rolls exhibit a rather
large wavelength of ;625 m in the southern zone. The
roll wavelength difference between the northern and
southern zones is ;225 m, which is much larger than that
in the lidar observations and in the simulation from
DS3_DA.
c. Near-surface streaks of wind speed and their
relation to the HCRs
Because the HCRs develop in a shallow sea-breeze
layer, their relationship with the near-surface winds is
examined and verified in this section. Figure 9a shows
the radial velocity of the PPI scan from the ENRI lidar
at a low elevation angle. Clearly, the streaky wind per-
turbations are embedded in the mean flow, with the axis
oriented in the downwind direction. The velocity dif-
ference between the streaks of fast-moving and slow-
moving flows is ;2 m s21. Comparing Figs. 9a and 8a, the
velocity streaks roughly represent the perturbations of
the major streamwise flow, given that the cross stream-
flow is secondary and slightly displaces the streak loca-
tions (Oda et al. 2010). Figure 9a shows that low-speed
streaks are well collocated with roll updrafts, while high-
speed streaks are located at the interval between up-
drafts. Such a good relation between streaks and rolls is
due to the wind perturbations that distinctly manifest
the convergence–divergence pattern as revealed by Iwai
et al. (2008). On the other hand, Fig. 9a presents a much
wider scan than the dual-Doppler lidar analysis domain
of Fig. 8a. Specifically, the streaks are weak and less
organized in the coastal areas in the southeast. The
streaks become well defined at Sendai Airport and over
inland areas in the northwest. Given the close link to
updrafts, these near-surface streaks indicate coastal
initiation and inland growth of sea-breeze HCRs.
Considering the streak-roll relation, the simulation of
near-surface winds is a key issue for realistically mod-
eling HCRs. Figure 9b shows the radial velocity derived
from the DS3_DA simulation. Low-speed streaks are
distinctly embedded in the sea breeze. These streaks
experience inland growth from the coast and become
evident over Sendai Airport. These features agree with
those observed by the PPI scan from the ENRI lidar
(cf. Figs. 9a and 9b). Figure 9b clearly shows that the
low-speed (high-speed) streaks are well collocated with
the roll updrafts (downdrafts). The DS3_DA simulation
seems to reproduce the near-surface streaks associated FIG. 9. Radial velocity with respect to the ENRI lidar over
with roll-scale convection. For comparison, Fig. 9c Sendai Airport. (a) PPI scan of the ENRI lidar at an elevation
angle of 18 (22.0–63.6 m AGL) at 1302 JST 19 Jun 2007, (b) CFD
shows that the simulated streaks notably change from
simulation from DS3_DA at 50 m AGL at 1305 JST, and (c) CFD
those in Figs. 9a and 9b. The streaks tend to stretch much simulation from DS3_noDA at 50 m AGL at 1305 JST. In (a), the
farther to the north, following a change in the wind di- solid lines mark the axis of the observed roll updrafts similarly to
rection in DS3_noDA. The spacing between the streaks Fig. 8a, and the dashed line represents the runway of Sendai Air-
port. In (b) and (c), the contours of vertical motion (0.5 m s21)
mark the axis of the simulated roll updrafts.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1863

FIG. 10. (a),(b) The in-plane winds and potential temperature simulated by the CFD model in
a cross-roll section with respect to the NICT lidar at 1305 JST 19 Jun 2007. (right) The mean
profile of potential temperature is plotted, with a dashed line for the surface temperature. The
location of the cross section is marked by the dashed line AB in Fig. 8c.

in the southwestern quadrant becomes large (;625 m). simulated by the CFD model is similar to that observed
These features correspond to the deficiency of the roll by the dual-Doppler lidar analysis [Fig. 3d in Iwai et al.
orientation, location, and wavelength in DS3_noDA (2008)], except the maximum updrafts are at a slightly
(Figs. 8e,f). However, the simulated velocity fields in lower level. In addition to dynamic features, Figs. 10a
Figs. 9b and 9c are smoother than the observations in and 10b show that the convective updrafts grow in
Fig. 9a, indicating that less-developed small-scale tur- thermal plumes, with a temperature ;1 K higher than
bulence occurs in the CFD model. the ambient air. Compensation downdrafts develop in
the spaces between updrafts, where cool air mass can
d. Vertical structure of the HCRs and associated
nearly reach the surface and absorb sensible heat. As the
secondary circulation
heated surface flows converge toward convective zones,
Figure 10 portrays the vertical structure of HCRs in they act to refuel the updrafts. Such a manner, in which
a cross-roll plane at 1305 JST. The HCRs are indicated the thermally driven secondary circulation creates
by the presence of several counter-rotating circulations a feedback to support roll sustenance, is consistent with
in the internal boundary layer (IBL). The updrafts what has been reported for other types of roll convection
are strongest at ;80 m AGL, with a magnitude of (Etling and Brown 1993; Atkinson and Zhang 1996;
;1.5 m s21. The updrafts are closely associated with the Young et al. 2002; Liu et al. 2004; Miao and Chen 2008).
convergent (divergent) cross-roll flows in the lower Figure 10 shows that the CFD-simulated rolls extend
(upper) half of the IBL. Such a secondary circulation throughout the entire IBL. The inversion of the air

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1864 MONTHLY WEATHER REVIEW VOLUME 143

temperature, along with near-surface temperature,


suggests an IBL depth of 140–200 m in Fig. 10a and 150–
240 m in Fig. 10b over Sendai Airport. These values are
comparable to or slightly lower than the lidar observa-
tion of ;220 m (Iwai et al. 2008; Oda et al. 2010).
Figure 10a also shows that the temperature contour
exhibits a wavelike pattern at the IBL top and implies
the possible existence of gravity waves. A gravity wave
may be triggered by an overshooting updraft, while its
persistence helps modulate roll vortices (Kuettner et al.
1987; Liu and Sang 2011). The gravity wave is most ev-
ident over the airport runway, and it corresponds to the
regularly spaced rolls. Figure 10b portrays a slightly
warmer IBL and a weaker IBL-top inversion compared
with that in Fig. 10a. This difference may be due to
the more northward wind direction in DS3_noDA. The
marine air mass is expected to travel farther over the
heated land surface before arriving at Sendai Airport.
The well-mixed IBL seems to allow updrafts to develop
higher in Fig. 10b. In contrast to the rolls near the NICT
lidar site, the rolls exhibit relatively low heights in
the northeastern region, where the IBL is shallow near
the shore in the CFD simulations. Therefore, a good
relation between the roll height and IBL depth, as re-
ported by previous studies, is also observed in the sea-
breeze HCRs.
Combining the roll wavelength (;515 m) and IBL
depth (;220 m) near the NICT lidar site, the aspect ratio
of the observed rolls is estimated to be 2.34, as also in-
dicated in early analyses (Iwai et al. 2008; Oda et al.
2010). Correspondingly, the aspect ratios of the CFD-
simulated rolls are estimated to be 2.38–3.39 in DS3_DA
and 2.60–4.17 in DS3_noDA. These values are slightly
larger than the lidar observations but are rather small
compared with the typical values (ranging from 2 to 20)
of roll family (Atkinson and Zhang 1996; Young et al.
2002). Many previous studies that use satellite cloud
images have only counted the well-developed rolls with
clouds to estimate the roll wavelength (Rao et al. 1999).
In this particular case, the roll updrafts do not induce
cloud formation because they occur in a shallow sea-
breeze layer with relatively dry land conditions. Thus,
sea-breeze HCRs, while likely common in coastal areas
in summer, are not well detected by conventional ob-
servations because of their small wavelengths, shallow
structures, and lack of clouds.
To illustrate the detailed structure of HCRs, Fig. 11a
portrays the high-resolution variables measured by a heli-
copter flight over Sendai Airport. The west–east-oriented
flight path is 5450 m long, with a spatial resolution of
FIG. 11. Vertical motion, along-roll wind speed, and temperature
1.16 m. The flight height of ;95 m AGL is nearly in the perturbations in a west–east section over Sendai Airport at 95 m AGL.
middle of the sea-breeze layer, where the strongest up- (a) Observations from a helicopter flight for 112 s starting at 1310:50
drafts occur. Figure 11a shows that updrafts are active JST 19 Jun 2007, with the original (smoothed) series in thin (bold)
lines. The flight path is plotted in Fig. 8b. (b) CFD simulation from
DS3_DA and (c) CFD simulation from DS3_noDA at 1305 JST.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1865

FIG. 12. Energy spectrum estimated from the data contained in Fig. 11 at the helicopter flight
path. The dashed line denotes the K25/3 inertial range slope. The wavenumber has been
multiplied by the path distance; the wavelength has been projected to the cross-roll plane.

over the concrete surface of the airport runway, while smaller than the fluxes estimated from the helicopter
they are less pronounced over the rice paddy fields and data when filtered to roll mode. We also note that the
shore regions. The major updrafts are generally spaced CFD simulation reproduces smooth variables and tends
over several hundreds of meters; thus, the IBL convection to underestimate small-scale turbulence. An analysis
has a roll-scale mode. The updrafts are overlapped by of the turbulence spectra (Fig. 12) confirms that the
small-scale turbulence that spans tens of meters (the thin CFD model captures roll-scale energy peak but falters in
line in Fig. 11a). For a roll-scale mode (bold line), the the reproduction of turbulence at an inertial range of
small-scale turbulence is filtered out by a running window ;100 m or smaller, despite some improvement com-
of 100 m. Figure 11a shows that roll updrafts (downdrafts) pared to mesoscale model. A possible reason is that the
coincide with the perturbations of low speed (high speed) transition from smooth mesoscale to resolved turbulence
and high temperature (low temperature), with a correla- in a LES model does not guarantee rapid growth of tur-
tion coefficient of 20.37 and 10.69, respectively. Based bulence in the nested LES area (Mirocha et al. 2014;
on the eddy correlation method, the roll-scale momen- Muñoz-Esparza et al. 2014). The relatively stable inflow
tum and heat fluxes are estimated as 20.148 m2 s22 and may also hinder high-frequency turbulence over the
0.134 K m s21, respectively. These fluxes account for upstream ocean surface in the LES domain. Other causes
two-thirds of the total fluxes estimated from the per- involve the numerical dissipation sources from the subgrid
turbation quantities (thin lines in Fig. 11a when re- model in LES and the diffusion in the advection scheme
moving the background value of linear regression). (Lilly 1962; Smagorinsky 1963; Beare 2014). Although this
Therefore, roll convection may be effective in trans- study focuses on roll-scale convection, it calls for im-
porting momentum and heat vertically within the IBL. proved nesting strategies in the future to reproduce small-
Figure 11b shows that the CFD simulation from scale turbulent flows under actual weather conditions.
DS3_DA appears to capture the major mode of the rolls. Figure 11c shows that the cross section from DS3_noDA
The simulated updrafts are evident over the airport runway shares many common features with that from DS3_DA.
but less pronounced over the rice paddy fields, which are DS3_noDA generally reproduces the roll-scale per-
consistent with helicopter measurements. The updrafts/ turbations, the relation between the updrafts and
downdrafts also concur with the fluctuations of the wind warm/low-speed air, and the regional differences over
speed and temperature. The resolved vertical transport inhomogeneous surfaces. We note in Fig. 11c that the
of momentum and heat are estimated as 20.085 m2 s22 amplitudes of the major updrafts, wind speed, and
and 0.077 K m s21, respectively. These values are somewhat temperature fluctuations are slightly larger than those in

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1866 MONTHLY WEATHER REVIEW VOLUME 143

Fig. 11b. The spacing between the updrafts over the kinetic buoyancy flux; and u0 w0 and y 0 w0 are the along- and
airport runway is also larger than that in Fig. 11b. These cross-roll kinetic momentum fluxes, respectively. The first
features manifest stronger rolls that develop in a well- term on the right is the TKE production by buoyancy, and
mixed IBL in DS3_noDA. The resolved vertical trans- it represents the thermal forcing on rolls. The second term
port of momentum and sensible heat are 20.095 m2 s22 is the TKE production by wind shear, and it expresses the
and 0.092 K m s21, respectively. These values are slightly shear forcing on rolls. The third (fourth) term is the TKE
larger than those in DS3_DA and are more similar to the change induced by turbulence (pressure) transport. These
observed roll-scale fluxes. two transport terms only redistribute TKE vertically and
On the other hand, the helicopter measurements in- do not contribute to the TKE source; D is the TKE loss by
dicate that most of the major updrafts have a maximum viscous dissipation. The TKE production terms are dis-
speed of more than 2 m s21, while the downdrafts are cussed here as we focus on roll formation mechanisms
approximately 21 m s21 (Fig. 11a). The updrafts are (Weckwerth et al. 1997), whereas the effects of all of the
confined to local sharp peaks, while the weak downdrafts production and transport terms on the vertical growth of
are more extensive; the updrafts and downdrafts are thus the rolls are investigated in Part II of this study.
asymmetric (Deardorff et al. 1969; Raasch and Harbusch To estimate the observed TKE budget terms, the mo-
2001). The updraft peaks are clearly coincident with the mentum and heat fluxes are derived using either lidar-
narrow bands of low wind speed and high temperature retrieved winds or helicopter measurements; the IBL
perturbations. Note that these features are not identified wind shear is given by the lidar-retrieved wind profile.
in the lidar-retrieval data at an effectively lower resolu- The shear production term derived from the lidar data is
tion (Figs. 8a,b). Figures 11b and 11c show that the sim- valid from 25 to 175 m AGL. The shear and buoyancy
ulated updrafts are more intense and narrower than the production terms are also derived from three helicopter
downdrafts; this finding is consistent with the helicopter flights along a west–east route over the airport runway at
measurements. Such asymmetric features are also ob- 45, 95, and 145 m AGL. The filtered helicopter data are
served in the other CFD simulations at a high resolution used in this TKE analysis to consider the roll-scale mode
(Liu et al. 2004). These features are closely related to the of the perturbations. Because the time lag between flights
narrow convergent zones, as indicated by the low-speed is ;6 min, we assumed that the change in mesoscale en-
streaks (Figs. 9b,c). The convective updrafts develop in vironment is small during the helicopter measurements.
the narrow convergent zones of the thermally unstable Figure 13a shows that, in the observation analysis, the
air, while the downdrafts behave like passive compensa- roll-energy production by shear is visible throughout the
tion flows (Fig. 10). The CFD model appears to re- IBL, with a maximum near the surface. The roll pro-
produce the intense/asymmetric roll-scale updrafts, duction by buoyancy is clearly observed in the middle of
despite the originally weak/uniform rolls in the absence the IBL. Figures 13b and 13c show that the budget profiles
of resolved smaller turbulence. derived from the two CFD runs exhibit magnitudes and
vertical structures that are similar to those in Fig. 13a,
except that the level of maximum buoyancy is relatively
5. Forcing mechanisms of the sea-breeze HCRs lower. This similarity indicates that the CFD model well
reproduce the roll dynamics. Figures 13a–c illustrate that
To illustrate roll-forcing mechanisms, we estimate the
shear and buoyancy effects are comparable in producing
production terms of the resolved roll-scale turbulent
roll kinetic energy near the surface, while buoyancy
kinetic energy (TKE) budget. Following LeMone (1973)
dominates most of the IBL above 40 m AGL. Such a shear
and Moeng and Sullivan (1994), the terms of the roll-
effect in roll convection is distinct from the unorganized
scale TKE budget are given by
or cellular convection in which buoyancy dominates both
  near the surface and throughout the CBL (Moeng and
› 0 g 0 0 ›U ›V ›w0 E0
E 5 uy w 2 u0 w0 1 y 0 w0 2 Sullivan 1994; Weckwerth et al. 1997). Here the evidence
›t uy ›z ›z ›z
from the observations and simulations consistently sug-
1 ›w0 p0 gest that a combination of dynamic and thermal effects
2 2 D, (1)
r0 ›z acts to drive and sustain roll convection in the sea breeze.
Figures 13b and 13c show that an IBL integration of
where the primes denote roll-scale perturbations and the the TKE production by buoyancy is approximately 2–3
overbars denote regional means. The roll-scale pertur- times greater than that produced by shear. Thus, buoy-
bations, as a kind of low-frequency turbulence, can be ancy serves a major energy source of convection: it
defined as the deviations from the spatially averaged manifests an evident thermal effect induced by the cool
means (Weckwerth et al. 1997). The term u0y w0 is the air intrusion over the heated surface to drive CBL

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1867

FIG. 13. Vertical profiles of the TKE production by buoyancy and shear over Sendai Airport. (a) Lidar (square with lines) and helicopter
(square only) measurements during 1300–1318 JST 19 Jun 2007, (b) CFD simulation from DS3_DA, and (c) CFD simulation from
DS3_noDA. The triangles mark the mean IBL height, which is estimated either from the NICT lidar observation of the signal-to-noise
ratio gradient (Iwai et al. 2008) or from the CFD-simulated inversion near the NICT lidar site (Fig. 10).

convection. The maximum buoyancy is located at ;60 m AGL and is distinct from a near-zero value in Fig. 13b.
AGL, where roll updrafts significantly strengthen; the This extra buoyancy might explain the higher extent of the
buoyancy decreases by half at ;100 m AGL. With the roll updrafts in DS3_noDA. Buoyancy due to thermal
diminishing buoyancy aloft, the roll updrafts weaken with instability greatly determines whether any convection will
increasing height, as shown in Fig. 10a. The buoyancy occur (Etling and Brown 1993; Atkinson and Zhang
reduces to near zero at the IBL top and above, where the 1996). Here, thermal instability also has a strong influence
updrafts dissipate. A comparison of Figs. 13b and 13c on the intensity and vertical extent of roll updrafts.
shows that the maximum buoyancy in DS3_noDA is 10% Figures 13a–c show that the shear production of roll
larger than that in DS3_DA. These values correspond to energy is most evident near the surface. The maximum
slightly stronger updrafts in DS3_noDA, as shown in production is mainly due to the near-surface shear
Figs. 10b and 11c. Another feature of Fig. 13c is that (likely related to surface roughness) that is several times
buoyancy is still observable at the levels of 150–230 m larger than the IBL shear (Figs. 14a–c). The profile is

FIG. 14. As in Fig. 13, but for the wind speed, momentum fluxes, and heat fluxes.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1868 MONTHLY WEATHER REVIEW VOLUME 143

similar to that reported by other analyses, suggesting parameter 2zi/L is estimated as 1.31–1.87 (1.46–2.38)
that the low-level shear is more important than the CBL over Sendai Airport. The values are comparable to that
shear in producing roll energy (Pennell and LeMone derived from lidar and helicopter measurements (1.79)
1974; Kristovich 1993; Weckwerth et al. 1997). Wind (Iwai et al. 2008). Both the observation and simulation
shear also serves to organize convection into along- analyses manifest a moderately unstable CBL that fa-
shear sheets and to suppress modes that are transverse vors the existence of roll convection (Deardorff 1972). A
to it (e.g., Kuo 1963; Asai 1964, 1970). In the case of the relatively large instability in DS3_noDA is consistent
sea breeze, the role of wind shear in maintaining roll with the increased buoyancy-to-shear ratio in the roll-
structures is most efficient near the surface, as shown in energy production (Fig. 13c). This enhancement corre-
Figs. 13a–c. This effect is thought to accumulate warm sponds to the roll aspect ratio that increases from 2.38–
fluid into low-speed streaks at the lowest level, localizing 3.39 in DS3_DA to 2.60–4.17 in DS3_noDA. The results
buoyancy forces to sustain convective bands (Lin et al. agree with previous studies that report rolls with larger
1997; Khanna and Brasseur 1998). Therefore, the dy- aspect ratios developing with larger 2zi/L parameters
namic effect due to near-surface shear plays a key role in (e.g., Kuo 1963; Kelly 1984; Weckwerth et al. 1997). We
organizing and maintaining the onshore HCRs within also note that DS3_DA, compared with DS3_noDA,
the sea breeze. Such shear and buoyancy forcings, while reproduces a more realistic roll aspect ratio observed
experiencing a gradual inland evolution, also generally over the airport (2.34). Therefore, capturing the CBL
hold for inland rolls, which are further discussed in structure may be related to roll-forcing mechanisms,
Part II of this study. which contribute to realistic simulations of roll features.
Figures 13a–c show that the shear production of roll
energy occurs above 40 m AGL and helps sustain the roll
structure throughout the IBL. Note that the shear pro- 6. Conclusions and future perspectives
duction term consists of wind shear and momentum
Forecasting sea-breeze HCRs over coastal cities is
fluxes. This production of roll energy in the IBL mostly
a challenge. We addressed this issue using an advanced
arises from momentum flux, as the IBL shear is as low as
numerical prediction system that integrates a mesoscale
;4 3 1023 s21 in the sea breeze (Figs. 14b,c). As re-
model with convective-scale data assimilation and
vealed previously, the interaction between rolls and
a building-resolving CFD model. We performed exper-
streaks is primarily responsible for the strong momen-
iments on a typical sea-breeze event in northeastern
tum flux. The roll–streak interaction thus contributes to
Japan, and we validated the numerical results using in-
the production of roll energy, even under a low-shear
tensive observations around Sendai Airport. The results
sea breeze. These findings support an early analysis of
indicate that a realistic simulation of HCRs can be
Weckwerth et al. (1997): although wind shear is neces-
achieved by combining reliable mesoscale weather
sary for roll convection, rolls may occur in a very low
forecasts with explicitly resolved individual rolls. The
CBL shear condition. We note that coherent downdrafts
findings are summarized as follows:
persisting in the IBL are crucial for transporting high
momentum from the IBL top to the surface (Fig. 14). 1) A nested LETKF system is applied to perform data
Because the downward momentum transport tends to assimilation experiments. The mesoscale-analyzed
increase wind speed in the downdraft zone, it provides data adequately capture the sea-breeze progress
a continuous kinetic source for sustaining the streaky during the daytime on 19 June 2007. Consistent with
flow patterns near the surface (i.e., Foster et al. 2006). surface records, the sea breeze exhibits a southeast-
In this respect, the IBL momentum fluxes and near- erly surge of 4–6 m s21, it penetrates inland with
surface streaks may work together to maintain roll-type a speed of 10–12 km h21, and it induces extensive
convection. cooling of ;218C h21. Compared with a simple
Just as in many theoretical and observational works, downscale forecast without data assimilation, the
roll formation involves the combined effect of dynamic analyzed data more accurately capture the passage
and thermal instabilities. The CBL instability parameter of sea-breeze front and consequent cooling at most
2zi/L, where zi is the CBL depth and L is the Obukhov observational sites. In particular, the analyzed data
length, is often used to measure the roles of dynamic and well represent the temporal variations in the surface
thermal instabilities in determining roll shape. Here L is temperature and low-level winds over Sendai Air-
calculated from the surface fluxes as in Stull (1988) and port. The data assimilation scheme thus provides an
Weckwerth et al. (1997), which has a value of 2107 m improved representation of the mesoscale weather
in DS3_DA and 2103 m in DS3_noDA. With zi within associated with the sea-breeze event in this particular
140–200 m in DS3_DA (150–240 m in DS3_noDA), the case study.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1869

2) The characteristics and structure of the HCRs simu- In this study, HCRs are strongly dependent on com-
lated by the nested CFD model are verified using dual- plex surfaces, as also noted in Iwai et al. (2008). The rolls
Doppler lidar and helicopter measurements over over Sendai Airport thus present a good opportunity to
Sendai Airport. The simulated rolls form along coastal study the influence of land use/buildings on roll structure
areas and develop over land. The rolls become active and evolution. Other studies also reported that roll
over buildup areas but are less pronounced over rice convection tends to occur over urban areas, where the
paddy fields; thus, they clearly respond to inhomoge- surface conditions are distinct from the surrounding
neous surfaces. The rolls are oriented from southeast areas (Kropfli and Kohn 1978; Newsom et al. 2008; Miao
to northwest, parallel to the CBL wind direction and and Chen 2008; Inagaki and Kanda 2010; Oda et al.
shear vectors. The rolls exhibit a wavelength of 400– 2012). Complex surfaces are known to involve thermal/
500 m and extend through the entire IBL with a depth dynamic effects that greatly regulate turbulent struc-
of ;200 m. The intense updrafts are typically rooted in tures and local circulations (Kanda et al. 2004; Prabha
the narrow bands of low-speed streaks and high- et al. 2007; Ashie and Kono 2011; Castillo et al. 2011;
temperature perturbations near the ground, while Kang and Bryan 2011; Inagaki et al. 2012; Park and Baik
the downdrafts are collocated with the high-speed 2013; Maronga and Raasch 2013). Additional effort is
streaks and low-temperature perturbations. The up- required to quantify the detailed impacts of complex
drafts (downdrafts) are associated with cross-roll surfaces on roll behaviors. Using explicit resolutions of
secondary flows that are convergent (divergent) in individual buildings, the CFD model can serve as a use-
the low levels of the IBL. These simulated HCRs ful tool to describe these effects. In Part II of this study,
better agree with lidar and helicopter observations we will perform additional experiments to illustrate how
than those in the experiment without the data assim- the sea-breeze HCRs respond to land-use heterogeneity
ilation. Even the specific locations of individual rolls and building morphology under actual weather condi-
can be reproduced with reasonably good precision. To tions (Part II).
our knowledge, this research produces the first re- Because the mesoscale model output is fed into the
alistic simulation of observed sea-breeze HCRs with CFD domain, the forecast error may introduce un-
a high accuracy and super high resolution over an certainty. This study produces an improved forecast
urban-scale domain. through a downscaling framework with convective-scale
3) The TKE production is estimated to elucidate the data assimilation. It is also noted that a further im-
roll-forcing mechanisms. The simulated budget pro- provement in the framework may involve two other is-
files are quite similar to those derived from the lidar sues: the ‘‘gray zone’’ and strategies to bridge mesoscale
and helicopter measurements; thus, the roll dynamics and microscale models. To improve the mesoscale
are well reproduced in the CFD model. Similarly to modeling at gray-zone resolutions, the parameterization
many theories of roll formation, wind shear and of the turbulent fluxes can be optimized using several
buoyancy forces work together to drive and sustain methods (Wyngaard 2004; Dorrestijn et al. 2013; Shin
roll convection in the sea breeze with moderate and Hong 2013). Beare (2014) recently proposed
instability and wind speed shear. Specifically, buoy- a length scale to quantify the effect of numerical dis-
ancy provides a major energy source for convection sipation at gray-zone resolutions. Ching et al. (2014)
occurrence and influences the strength and depth of suggested options for handling the convectively in-
updrafts. Buoyancy reflects the intrusion of cool duced secondary circulations that are poorly resolved
marine air over warm land that produces thermal in mesoscale models. On the other hand, downscaling
instability driving convection. The dynamic effects of from smooth mesoscale flows does not necessarily lead
wind shear and momentum fluxes also produce a large to rapid growth of turbulent flows in the nested LES
amount of roll-scale kinetic energy, particularly near domain. Mirocha et al. (2014) noted that nesting of
the surface. This process greatly regulates the streaky a finer LES within a coarse LES helps accelerate tur-
pattern of the winds and temperature at the lowest bulence growth. Muñoz-Esparza et al. (2014) de-
level; thus, it plays a key role in maintaining roll-type veloped a new method based upon the use of potential
convection. Considering that the IBL structure and temperature perturbations near the inflow boundaries
surface conditions closely relate to roll-forcing phys- in the LES domain that substantially accelerates the
ics, their adequate representation in the CFD model development of turbulence at a low computational
explains the improved forecast of the sea-breeze cost. Their idealized experiments also indicate that the
HCRs. The application of high-precision downscal- turbulence growth in a nested LES depends on
ing may lead to the more realistic HCRs simulations boundary layer stability and surface geometry. These
in this study, in contrast to previous idealized studies. results motivate us to evaluate and improve the

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1870 MONTHLY WEATHER REVIEW VOLUME 143

downscaling framework for a wide range of realistic Atkinson, B. W., and J. W. Zhang, 1996: Mesoscale shallow con-
weather and surface conditions. vection in the atmosphere. Rev. Geophys., 34, 403–431,
doi:10.1029/96RG02623.
This study is one part of the Strategic Programs for
Baik, J.-J., S.-B. Park, and J.-J. Kim, 2009: Urban flow and dis-
Innovative Research (SPIRE) of Japan that aims to persion simulation using a CFD model coupled to a mesoscale
improve high-precision mesoscale forecasts (Saito et al. model. J. Appl. Meteor. Climatol., 48, 1667–1681, doi:10.1175/
2013). Here we evaluated the performance of a down- 2009JAMC2066.1.
scaling system for one typical weather case. Because the Beare, R. J., 2014: A length scale defining partially-resolved
boundary-layer turbulence simulations. Bound.-Layer
sea breeze is a common phenomenon in summer, veri-
Meteor., 151, 39–55, doi:10.1007/s10546-013-9881-3.
fication should be conducted for more cases to estimate Castillo, M. C., A. Inagaki, and M. Kanda, 2011: The effects of
the extent to which we can improve local weather inner- and outer-layer turbulence in a convective boundary
forecasts by better modeling HCRs. In addition to sea- layer on the near-neutral inertial sublayer over an urban-like
breeze events, the nested system of data assimilation is surface. Bound.-Layer Meteor., 140, 453–469, doi:10.1007/
a powerful tool to improve forecast accuracy for a vari- s10546-011-9614-4.
Chen, G., X. Zhu, W. Sha, T. Iwasaki, H. Seko, K. Saito, H. Iwai,
ety of mesoscale weather, including extreme events such and S. Ishii, 2015: Toward improved forecasts of sea-breeze
as heavy rainfalls and typhoons (Seko et al. 2011, 2013). horizontal convective rolls at super high resolutions. Part II:
Here a combination of reliable mesoscale forecasts and The impacts of land use and buildings. Mon. Wea. Rev., 143,
the CFD model is feasible for the super-high-resolution 1885–1906, doi:10.1175/MWR-D-14-00230.1.
modeling of regional weather over complex surfaces. Ching, J., R. Rotunno, M. LeMone, A. Martilli, B. Kosovic, P. A.
Jimenez, and J. Dudhia, 2014: Convectively induced second-
One of the most valuable applications of this concept is
ary circulations in fine-grid mesoscale numerical weather
to simulate mesoscale weather in large cities with many prediction models. Mon. Wea. Rev., 142, 3284–3302,
tall buildings. Because DS3 can explicitly resolve doi:10.1175/MWR-D-13-00318.1.
buildings, it is potentially applicable for realistic mod- Deardorff, J. W., 1972: Numerical investigation of neutral and
eling of finescale winds and temperature within a city. unstable planetary boundary layers. J. Atmos. Sci., 29, 91–115,
doi:10.1175/1520-0469(1972)029,0091:NIONAU.2.0.CO;2.
Further verification of DS3 or similar downscaling sys-
——, G. E. Willis, and D. K. Lilly, 1969: Laboratory investigation
tems is warranted to improve the forecast skill regarding of non-steady penetrative convection. J. Fluid Mech., 35, 7–31,
urban weather and to establish operational numerical doi:10.1017/S0022112069000942.
prediction at a resolution of a few meters in the future. Dorrestijn, J., D. T. Crommelin, A. P. Siebesma, and H. J. J. Jonker,
2013: Stochastic parameterization of shallow cumulus convec-
tion estimated from high-resolution model data. Theor. Com-
Acknowledgments. The authors thank the three
put. Fluid Dyn., 27, 133–148, doi:10.1007/s00162-012-0281-y.
anonymous reviewers for their helpful comments and Eito, H., M. Murakami, C. Muroi, T. Kato, S. Hayashi, H. Kuroiwa,
constructive suggestions. This study was supported by and M. Yoshizaki, 2010: The structure and formation mecha-
the Strategic Programs for Innovative Research (SPIRE) nism of transversal cloud bands associated with the Japan-Sea
funded by the Japan Ministry of Education, Culture, polar-airmass convergence zone. J. Meteor. Soc. Japan, 88,
Sports, Science and Technology (MEXT). The numerical 625–648, doi:10.2151/jmsj.2010-402.
Etling, D., and R. A. Brown, 1993: Roll vortices in the planetary
calculations were performed using the K computer at boundary layer: A review. Bound.-Layer Meteor., 65, 215–248,
RIKEN Advanced Institute for Computational Science doi:10.1007/BF00705527.
(Proposals hp120282, hp130012, and hp140220) and the Foster, R. C., F. Vianey, P. Drobinski, and P. Carlotti, 2006: Near-
supercomputing resource SX-9 at the Cyberscience surface coherent structures and the vertical momentum flux in
Center, Tohoku University. The observational data were a large-eddy simulation of the neutrally-stratified boundary
layer. Bound.-Layer Meteor., 120, 229–255, doi:10.1007/
collected in a field experiment funded by the MEXT (PI:
s10546-006-9054-8.
Prof. T. Iwasaki, 19204046) through the collaboration of Gryschka, M., and S. Raasch, 2005: Roll convection during a cold air
NICT, ENRI, JAXA, and Tohoku University. outbreak: A large eddy simulation with stationary model domain.
Geophys. Res. Lett., 32, L14805, doi:10.1029/2005GL022872.
Hayase, T., J. A. C. Humphrey, and R. Greif, 1992: A consistently
REFERENCES formulated QUICK scheme for fast and stable convergence
using finite volume iterative calculation procedures. J. Com-
Asai, T., 1964: Cumulus convection in the atmosphere with vertical put. Phys., 98, 108–118, doi:10.1016/0021-9991(92)90177-Z.
wind shear: Numerical experiment. J. Meteor. Soc. Japan, 42, Inagaki, A., and M. Kanda, 2010: Organized structure of active
245–259. turbulence over an array of cubes within the logarithmic
——, 1970: Stability vertical of a plane shear parallel and flow with layer of atmospheric flow. Bound.-Layer Meteor., 135, 209–
variable unstable stratification. J. Meteor. Soc. Japan, 48, 228, doi:10.1007/s10546-010-9477-0.
129–139. ——, M. C. L. Castillo, Y. Yamashita, M. Kanda, and H. Takimoto,
Ashie, Y., and T. Kono, 2011: Urban-scale CFD analysis in support 2012: Large-eddy simulation of coherent flow structures
of a climate-sensitive design for the Tokyo Bay area. Int. within a cubical canopy. Bound.-Layer Meteor., 142, 207–222,
J. Climatol., 31, 174–188, doi:10.1002/joc.2226. doi:10.1007/s10546-011-9671-8.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


MAY 2015 CHEN ET AL. 1871

Ishii, S., and Coauthors, 2007: Temporal evolution and spatial Maronga, B., and S. Raasch, 2013: Large-eddy simulations of sur-
structure of the local easterly wind ‘‘Kiyokawa-dashi’’ in face heterogeneity effects on the convective boundary layer
Japan. Part I: Coherent Doppler lidar observations. J. Meteor. during the LITFASS-2003 Experiment. Bound.-Layer Me-
Soc. Japan, 85, 797–813, doi:10.2151/jmsj.85.797. teor., 146, 17–44, doi:10.1007/s10546-012-9748-z.
Iwai, H., and Coauthors, 2008: Dual-Doppler lidar observation of Matayoshi, N., H. Inokuchi, K. Yazawa, and Y. Okuno, 2005:
horizontal convective rolls and near-surface streaks. Geophys. Development of airborne ultrasonic velocimeter and its ap-
Res. Lett., 35, L14808, doi:10.1029/2008GL034571. plication to helicopters. AIAA Atmospheric Flight Mechanics
Kanda, M., R. Moriwaki, and F. Kasamatsu, 2004: Large eddy Conf. and Exhibit, San Francisco, CA, AIAA, AIAA 2005-
simulation of turbulent organized structure within and above 6118. [Available online at http://arc.aiaa.org/doi/abs/10.2514/
explicitly resolved cubic arrays. Bound.-Layer Meteor., 112, 6.2005-6118.]
343–368, doi:10.1023/B:BOUN.0000027909.40439.7c. Miao, S., and F. Chen, 2008: Formation of horizontal convective
Kang, S.-L., and G. H. Bryan, 2011: A large-eddy simulation study of rolls in urban areas. Atmos. Res., 89, 298–304, doi:10.1016/
moist convection initiation over heterogeneous surface fluxes. j.atmosres.2008.02.013.
Mon. Wea. Rev., 139, 2901–2917, doi:10.1175/MWR-D-10-05037.1. Miller, S. T. K., B. D. Keim, R. W. Talbot, and H. Mao, 2003: Sea
Kelly, R. D., 1984: Horizontal roll and boundary-layer interrelation- breeze: Structure, forecasting, and impacts. Rev. Geophys.,
ships observed over Lake Michigan. J. Atmos. Sci., 41, 1816–1826, 41, 1011, doi:10.1029/2003RG000124.
doi:10.1175/1520-0469(1984)041,1816:HRABLI.2.0.CO;2. Mirocha, J., B. Kosovic, and G. Kirkil, 2014: Resolved turbulence
Khanna, S., and J. G. Brasseur, 1998: Three-dimensional buoy- characteristics in large-eddy simulations nested within mesoscale
ancy- and shear-induced local structure of the atmospheric simulations using the Weather Research and Forecasting Model.
boundary layer. J. Atmos. Sci., 55, 710–743, doi:10.1175/ Mon. Wea. Rev., 142, 806–831, doi:10.1175/MWR-D-13-00064.1.
1520-0469(1998)055,0710:TDBASI.2.0.CO;2. Miyoshi, T., and K. Aranami, 2006: Applying a four-dimensional
Komatsubara, T., and N. Kaku, 2005: The status of wake vortex Local Ensemble Transform Kalman Filter (4D-LETKF) to
research in Japan. Proc. 13th Coherent Laser Radar Conf., the JMA Nonhydrostatic Model (NHM). SOLA, 2, 128–131,
Kamakura, Japan, National Institute of Information and doi:10.2151/sola.2006-033.
Communications Technology, 140–143. Moeng, C.-H., 1986: Large-eddy simulation of a stratus-topped
Kristovich, D. A. R., 1993: Mean circulations of boundary-layer boundary layer. Part I: Structure and budgets. J. Atmos.
rolls in lake-effect snow storms. Bound.-Layer Meteor., 63, Sci., 43, 2886–2900, doi:10.1175/1520-0469(1986)043,2886:
293–315, doi:10.1007/BF00710463. LESOAS.2.0.CO;2.
Kropfli, R. A., and N. M. Kohn, 1978: Persistent horizontal ——, and P. P. Sullivan, 1994: A comparison of shear- and
rolls in the urban mixed layer as revealed by dual-Doppler buoyancy-driven planetary boundary layer flows. J. Atmos.
radar. J. Appl. Meteor., 17, 669–676, doi:10.1175/ Sci., 51, 999–1022, doi:10.1175/1520-0469(1994)051,0999:
1520-0450(1978)017,0669:PHRITU.2.0.CO;2. ACOSAB.2.0.CO;2.
Kuettner, J. P., P. A. Hildebrand, and T. L. Clark, 1987: Convection Muñoz-Esparza, D., B. Kosovic, J. Mirocha, and J. van Beeck,
waves: Observations of gravity wave systems over con- 2014: Bridging the transition from mesoscale to microscale
vectively active boundary layers. Quart. J. Roy. Meteor. Soc., turbulence in numerical weather prediction models. Bound.-
113, 445–467, doi:10.1002/qj.49711347603. Layer Meteor., 153, 409–440, doi:10.1007/s10546-014-9956-9.
Kuo, H. L., 1963: Perturbations of plane couette flow in strati- Newsom, R. K., R. Calhoun, D. Ligon, and J. Allwine, 2008: Lin-
fied fluid and origin of cloud streets. Phys. Fluids, 6, 195– early organized turbulence structures observed over a sub-
211, doi:10.1063/1.1706719. urban area by dual-Doppler lidar. Bound.-Layer Meteor., 127,
LeMone, M. A., 1973: The structure and dynamics of horizontal 111–130, doi:10.1007/s10546-007-9243-0.
roll vortices in the planetary boundary layer. J. Atmos. Oda, R., and Coauthors, 2010: Doppler lidar observations of the
Sci., 30, 1077–1091, doi:10.1175/1520-0469(1973)030,1077: coherent structures in the internal boundary layer. Ninth
TSADOH.2.0.CO;2. Symp. on the Urban Environment, Keystone, CO, Amer.
——, 1976: Modulation of turbulence energy by longitudinal rolls in an Meteor. Soc., 10.1. [Available online at https://ams.confex.
unstable planetary boundary layer. J. Atmos. Sci., 33, 1308–1320, com/ams/19Ag19BLT9Urban/techprogram/paper_172111.htm.]
doi:10.1175/1520-0469(1976)033,1308:MOTEBL.2.0.CO;2. ——, H. Iwai, A. Inagaki, S. Ishii, S. Satoh, S. Sekizawa, K. Mizutani,
Leonard, B. P., 1979: A stable and accurate convection modeling and Y. Murayama, 2012: Doppler lidar observation of turbulent
procedure based on quadratic upstream interpolation. mixing in the urban atmospheric boundary layer. Proc. Eighth
Comput. Methods Appl. Mech. Eng., 19, 59–98, doi:10.1016/ Int. Conf. on Urban Climates, Dublin, Ireland, University
0045-7825(79)90034-3. College Dublin, 295.
Lilly, D. K., 1962: On the numerical simulation of buoyant convection. Orlanski, I., 1976: A simple boundary condition for unbounded
Tellus, 14A, 148–172, doi:10.1111/j.2153-3490.1962.tb00128.x. hyperbolic flows. J. Comput. Phys., 21, 251–269, doi:10.1016/
Lin, C.-L., C.-H. Moeng, P. P. Sullivan, and J. C. McWilliams, 1997: 0021-9991(76)90023-1.
The effect of surface roughness on flow structures in a neu- Park, S.-B., and J.-J. Baik, 2013: A large-eddy simulation study of
trally stratified planetary boundary layer. Phys. Fluids, 9, thermal effects on turbulence coherent structures in and
3235–3249, doi:10.1063/1.869439. above a building array. J. Appl. Meteor. Climatol., 52, 1348–
Liu, A. Q., G. W. K. Moore, K. Tsuboki, and I. A. Renfrew, 2004: 1365, doi:10.1175/JAMC-D-12-0162.1.
A high-resolution simulation of convective roll clouds Patankar, S. V., 1980: Numerical Heat Transfer and Fluid Flow.
during a cold-air outbreak. Geophys. Res. Lett., 31, L03101, Taylor & Francis Press, 214 pp.
doi:10.1029/2003GL018530. Pennell, W. T., and M. A. LeMone, 1974: An experimental study
Liu, H., and J. Sang, 2011: Numerical simulation of roll vortices in of turbulence structure in the fair-weather trade wind
the convective boundary layer. Adv. Atmos. Sci., 28, 477–482, boundary layer. J. Atmos. Sci., 31, 1308–1323, doi:10.1175/
doi:10.1007/s00376-010-9229-6. 1520-0469(1974)031,1308:AESOTS.2.0.CO;2.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC


1872 MONTHLY WEATHER REVIEW VOLUME 143

Prabha, T. V., A. Karipot, and M. W. Binford, 2007: Characteristics Meteorological Research Note, Y. Fujiyoshi, Ed., Vol. 219,
of secondary circulations over an inhomogeneous surface Meteorological Society of Japan Press, 21–26.
simulated with large-eddy simulation. Bound.-Layer Meteor., ——, T. Kawamura, and H. Ueda, 1991: A numerical study on
123, 239–261, doi:10.1007/s10546-006-9137-6. sea/land breezes as a gravity current: Kelvin–Helmholtz bil-
Raasch, S., and G. Harbusch, 2001: An analysis of secondary cir- lows and inland penetration of the sea-breeze front. J. Atmos.
culations and their effects caused by small-scale surface in- Sci., 48, 1649–1665, doi:10.1175/1520-0469(1991)048,1649:
homogeneities using large-eddy simulation. Bound.-Layer ANSOSB.2.0.CO;2.
Meteor., 101, 31–59, doi:10.1023/A:1019297504109. Shin, H. H., and S.-Y. Hong, 2013: Analysis of resolved and pa-
Rao, P. A., H. E. Fuelberg, and K. K. Droegemeier, 1999: High- rameterized vertical transports in convective boundary layers
resolution modeling of the Cape Canaveral area land– at gray-zone resolutions. J. Atmos. Sci., 70, 3248–3261,
water circulations and associated features. Mon. Wea. Rev., doi:10.1175/JAS-D-12-0290.1.
127, 1808–1821, doi:10.1175/1520-0493(1999)127,1808: Simpson, J. E., 1969: A comparison between laboratory currents
HRMOTC.2.0.CO;2. and atmospheric density currents. Quart. J. Roy. Meteor. Soc.,
Saito, K., 2012: The JMA nonhydrostatic model and its applica- 95, 758–765, doi:10.1002/qj.49709540609.
tions to operation and research. Atmospheric Model Appli- Smagorinsky, J., 1963: General circulation experiments with the
cations, I. Yucel, Ed., InTech, 85–110, doi:10.5772/35368. primitive equations. Mon. Wea. Rev., 91, 99–164, doi:10.1175/
——, and Coauthors, 2006: The operational JMA nonhydrostatic 1520-0493(1963)091,0099:GCEWTP.2.3.CO;2.
mesoscale model. Mon. Wea. Rev., 134, 1266–1298, Stull, R. B., 1988: An Introduction to Boundary Layer Meteorology.
doi:10.1175/MWR3120.1. Kluwer Academic, 666 pp.
——, J. Ishida, K. Aranami, T. Hara, T. Segawa, M. Narita, and Weckwerth, T. M., J. W. Wilson, and R. M. Wakimoto, 1996:
Y. Honda, 2007: Nonhydrostatic atmospheric models and Thermodynamic variability within the convective boundary
operational development at JMA. J. Meteor. Soc. Japan, 85B, layer due to horizontal convective rolls. Mon. Wea. Rev.,
271–304, doi:10.2151/jmsj.85B.271. 124, 769–784, doi:10.1175/1520-0493(1996)124,0769:
——, and Coauthors, 2013: Super high-resolution mesoscale TVWTCB.2.0.CO;2.
weather prediction. J. Phys. Conf. Ser., 454, 012073, ——, ——, ——, and N. A. Crook, 1997: Horizontal convective
doi:10.1088/1742-6596/454/1/012073. rolls: Determining the environmental conditions support-
Seko, H., T. Miyoshi, Y. Shoji, and K. Saito, 2011: Data assimilation ing their existence and characteristics. Mon. Wea. Rev.,
experiments of precipitable water vapor using the LETKF 125, 505–526, doi:10.1175/1520-0493(1997)125,0505:
system: An intense rainfall event over Japan 28 July 2008. HCRDTE.2.0.CO;2.
Tellus, 63A, 402–412, doi:10.1111/j.1600-0870.2010.00508.x. Wyngaard, J. C., 2004: Towards numerical modeling in the
——, T. Tsuyuki, K. Saito, and T. Miyoshi, 2013: Development of ‘‘terra incognita.’’ J. Atmos. Sci., 61, 1816–1826, doi:10.1175/
a two-way nested LETKF system for cloud resolving model. 1520-0469(2004)061,1816:TNMITT.2.0.CO;2.
Data Assimilation for Atmospheric, Oceanic and Hydrologic ——, and C.-H. Moeng, 1992: Parameterizing turbulent diffusion
Applications II, S. K. Park and L. Xu, Eds., Springer-Verlag, through the joint probability density. Bound.-Layer Meteor.,
197–218. 60, 1–13, doi:10.1007/BF00122059.
Sha, W., 2002: Design of the dynamics core for a new-generation Young, G. S., D. A. R. Kristovich, M. R. Hjelmfelt, and R. C.
numerical model of the local meteorology (in Japanese). Foster, 2002: Rolls, streets, waves, and more: A review of
Kaiyo Mon., 2, 107–112. quasi-two-dimensional structures in the atmospheric bound-
——, 2008: Local meteorological model based on LES over the ary layer. Bull. Amer. Meteor. Soc., 83, 997–1001, doi:10.1175/
Cartesian coordinate and complex surface (in Japanese). 1520-0477(2002)083,0997:RSWAMA.2.3.CO;2.

Unauthenticated | Downloaded 12/09/20 06:41 PM UTC

You might also like