Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

VOLUME 49 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY DECEMBER 2010

A Numerical Study of Clear-Air Turbulence (CAT) Encounters over South


Korea on 2 April 2007

JUNG-HOON KIM AND HYE-YEONG CHUN


Department of Atmospheric Sciences, Yonsei University, Seoul, South Korea

(Manuscript received 20 November 2009, in final form 26 July 2010)

ABSTRACT

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


On 2 April 2007, nine cases of moderate-or-greater-level clear-air turbulence (CAT) were observed from
pilot reports over South Korea during the 6.5 h from 0200 to 0830 UTC. Those CAT events occurred in three
different regions of South Korea: the west coast, Jeju Island, and the eastern mountain areas. The charac-
teristics and possible mechanisms of the CAT events in the different regions are investigated using the
Weather Research and Forecasting model. The simulation consists of six nested domains focused on the
Korean Peninsula, with the finest horizontal grid spacing of 0.37 km. The simulated wind and temperature
fields in a 30-km coarse domain are in good agreement with those of the Regional Data Assimilation and
Prediction System (RDAPS) analysis data of the Korean Meteorological Administration and observed
soundings of operational radiosondes over South Korea. In synoptic features, an upper-level front associated
with strong meridional temperature gradients is intensified, and the jet stream passing through the central part
of the Korean Peninsula exceeds 70 m s21. Location and timing of the observed CAT events are reproduced
in the finest domains of the simulated results in three different regions. Generation mechanisms of the CAT
events revealed in the model results are somewhat different in the three regions. In the west coast area, the
tropopause is deeply folded down to about z 5 4 km because of the strengthening of an upper-level front, and
the maximized vertical wind shear below the jet core produces localized turbulence. In the Jeju Island area,
localized mixing and turbulence are generated on the anticyclonic shear side of the enhanced jet, where
inertial instability and ageostrophic flow are intensified in the lee side of the convective system. In the eastern
mountain area, large-amplitude gravity waves induced by complex terrain propagate vertically and sub-
sequently break down over the lee side of topography, causing localized turbulence. For most of the CAT
processes considered, except for the mountain-wave breaking, standard NWP resolutions of tens of kilo-
meters are adequate to capture the CAT events.

1. Introduction CAT forecasting skills and to assist pilots in avoiding


CAT, it is essential to investigate and understand pos-
Clear-air turbulence (CAT) is defined as the bumps of
sible mechanisms of individual CAT events through
in-flight aircraft directly induced by small-scale turbu-
numerical simulations (e.g., Clark et al. 2000; Lane et al.
lent eddies in free atmosphere within a cloud-free or
2003; Lane and Sharman 2008; Trier and Sharman 2009).
stratiform cloud condition (Ellrod et al. 2003). In the
A large portion of CAT has been explained by the
aviation industry, incidents of CAT have been recog-
Kelvin–Helmholtz instability (KHI), which can occur
nized as a serious concern because they occur unex-
when the local Richardson number (Rig 5 N2/VWS2,
pectedly at dominant flight altitudes of commercial
where N and VWS are the Brunt–Väisälä frequency and
aircraft, which result in problems such as occupant in-
vertical wind shear, respectively) is less than 0.25 in
juries, fuel losses, and flight delays. As air transportation
a stably stratified shear flow (Howard 1961; Miles 1961)
has increased, the importance of CAT forecasting has
and less than 1 in a nonlinear 3D flow (Abarbanel et al.
increased (Sharman et al. 2006). To develop more useful
1984, 1986; Miles 1986). Under this condition, turbu-
lent kinetic energy increases by extracting energy from
the mean flow and eventually cascades down to smaller-
Corresponding author address: Prof. Hye-Yeong Chun, Dept.
of Atmospheric Sciences, Yonsei University, 262 Seongsanno, scale eddies, which typically have scales similar to
Seodaemun-gu, Seoul 120-749, South Korea. aircraft (Sekioka 1970; Joseph et al. 2004). In the real
E-mail: chunhy@yonsei.ac.kr atmosphere, CAT events induced by KHI frequently

DOI: 10.1175/2010JAMC2449.1

Ó 2010 American Meteorological Society 2381


2382 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

occur above or below the jet stream where the VWS is strong vertical shear in the upper-level outflow of a
sufficiently strong to overcome its environmental stability. mesoscale convective system (MCS). The subsequent
In addition, the entrance region of the jet stream and the higher-resolution simulations of Trier et al. (2010) sug-
confluence zone of two different flows favor CAT gener- gested that this observed turbulence may have been re-
ation through upper-level frontogenesis (Mancuso and lated to shallow banded convection that was organized by
Endlich 1966; Roach 1970). The strong vertical wind shear the vertical shear in the near-neutral to weakly con-
and low static stability near the frontal zone result in KHI. vectively unstable outflow.
Ellrod and Knapp (1992) indicated that a turbulence index Another mechanism of CAT is a breakdown of gravity
1 (TI1 5 DEF 3 VWS, where DEF is total deformation) waves generated by complex terrain and convection. Jiang
that is based on upper-level frontogenesis is a good di- and Doyle (2004), Doyle et al. (2005), and Lane et al. (2009)
agnosis for CAT near the developing upper-level front and claimed that large-amplitude mountain waves and their
the jet core in a highly curved trough. However, it is not subsequent breaking over complex terrain are responsible
sufficient to describe a highly localized individual CAT for small-scale turbulence that directly affects aircraft.
event solely by using the KHI based on frontogenesis (e.g., Mountain waves can be a source of CAT in several ways.
Knox 1997; Kaplan et al. 2005; Ellrod and Knox 2010). First, the amplitude of the vertically propagating mountain

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


As improvement in observation techniques and com- waves increases with height because of decreasing air
puter capacity advances our understanding of CAT mech- density (Hines 1960; Lindzen 1967), which eventually leads
anisms, various other mechanisms of CAT have been to wave steepening, overturning, and subsequent breaking
suggested in addition to those related to the jet–front in the higher altitudes (Prusa et al. 1996; Doyle et al. 2005).
system. Inertial instability associated with anticyclonic Second, an increase in atmospheric stability near the tro-
flow has been considered as an important source of CAT, popause can reduce the vertical wavelength of mountain
although CAT events are less frequent in anticyclonic waves, which increases the potential for wave breaking
flow than in cyclonic flow. Knox (1997) showed that total (VanZandt and Fritts 1989). Third, stationary mountain
deformation induced by upper-level frontogenesis is not waves can break down when they approach a critical level
sufficient to explain CAT events in strong anticyclonic at which the background wind speed equals zero (Lilly
flow. Rather, inertial instability and/or geostrophic ad- 1978; Smith 1979, 1989; Wurtele et al. 1996; Smith et al.
justment could be regarded as a more plausible source 2002). With regard to convective gravity waves, Lane et al.
of CAT events in strong anticyclonic flow. Meanwhile, (2003) suggested that a severe CAT encounter near North
Clark et al. (2000) demonstrated that the location of the Dakota on 10 July 1997 was related to convectively induced
severe turbulence over Colorado on 9 December 1992 is gravity waves and their subsequent breaking above a well-
consistent with the region of observed and simulated developed convective cloud top. Lane and Sharman (2008)
horizontal vortex tubes (HVTs). In their hypothesis, also suggested that CAT induced by convectively generated
these well-mixed and rotating flows of the HVTs are gravity waves depends on background stability and VWS
produced when strong variations in the vertical compo- above a well-developed convective cloud.
nent of relative vorticity transit to the horizontal com- Climatologically, the region around the Korean Penin-
ponent of vorticity due to flow blocking or convergence sula and Japan has a stronger jet stream than any other
over the terrain. Combining the aforementioned con- region of the world during all seasons, even though the
cepts from Knox (1997) and Clark et al. (2000), Kaplan location of local maxima is different in each season (Koch
et al. (2005) suggested a forecasting index, the North et al. 2006). For this reason, potential CAT encounters
Carolina State University index 2 (NCSU2 5 $z 3 $M, diagnosed using several CAT indices are relatively high
where z and M are streamwise vertical vorticity and over East Asia, including over South Korea, as shown by
Montgomery streamfunction on an isentropic surface, Ellrod et al. (2003) and Jaeger and Sprenger (2007). In
respectively), and showed that the NCSU2 index is max- addition, the Korean Peninsula is covered mostly (more
imized in a region where the strong gradient of the than 70%) by complex structures of topography, as shown
streamwise vertical vorticity is perpendicular to the strong in Fig. 1, so that the prevailing westerly flows over South
pressure gradient force at an isentropic surface, which fi- Korea can produce mountain waves with a wide spectrum.
nally transits to the HVTs and causes localized turbulence. In these circumstances, the aforementioned generation
Kaplan et al. (2006) showed that the NCSU2 index can be mechanisms can be applied to the observed CAT events
a good indicator of severe turbulence in ageostrophi- over South Korea.
cally induced anticyclonic flow regions such as the lee side This study documents detailed investigations of
of well-developed convective systems. Trier and Sharman moderate or greater (MOG)-level CAT events over
(2009) reported that widespread turbulence indicated by South Korea on 2 April 2007. In the 6.5 h from 0200 to
in situ data from commercial aircraft was associated with 0830 UTC, 12 CAT events, including nine MOG-level
DECEMBER 2010 KIM AND CHUN 2383

TABLE 1. Time, latitude, longitude, flight level, and intensity of


nine MOG-level CAT events over South Korea reported in PIREPs
on 2 Apr 2007. Here, W, J, and E in the first column indicate the
three categorized regions: the west coast, Jeju Island, and the
eastern mountain areas of South Korea, respectively. Flight levels
(km) in the fifth column are calculated using the corresponding
standard atmospheric height AGL from the flight altitudes pro-
vided by conventional PIREPs.

Time Lat Lon Flight Turbulence


No. (UTC) (8N) (8E) level (km) intensity
1 (E) 0200 37.08 129.78 11.8 Moderate
2 (J) 0530 33.21 126.48 10.9 Severe
3 (W) 0600 36.80 126.69 7.2 Moderate
4 (W) 0700 37.78 125.70 5.8 Moderate
5 (E) 0730 36.20 129.23 12.2 Moderate
6 (W) 0750 37.58 126.17 5.0 Moderate
7 (E) 0800 36.86 129.46 12.6 Moderate

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


8 (J) 0810 32.43 125.73 10.8 Moderate
9 (J) 0830 33.61 126.98 9.8 Moderate

MOG-level CAT events that occurred during the 6.5 h


from 0200 (1100) to 0830 (1730) UTC (LST). Considering
that from 2003 to 2008 the average number of PIREPs and
FIG. 1. Locations of MOG-level CAT events over South Korea on
2 Apr 2007 superimposed on terrain height (shading). The numbers CAT encounters over South Korea has been only 3.4 and
of MOG-level CAT events correspond to those described in Table 1. 2 per day, respectively (Kim and Chun 2011), the occur-
rence of 12 PIREPs and 12 CAT events, with nine MOG-
level CAT events, in a single day is very unusual. The
CAT events, were reported in pilot reports (PIREPs)
atmospheric conditions over the Korean Peninsula during
over three different regions of South Korea: the west
this period might have been particularly vulnerable to gen-
coast, Jeju Island, and the eastern mountain areas. This
erate that much aircraft-scale turbulence. In the present
is the highest single-day frequency of CAT events over
study, we will exclusively investigate the nine MOG-level
South Korea for the last 10 yr (Kim and Chun 2008).
CAT events (CAT events hereafter refers to those nine
The Advanced Research Weather Research and Fore-
MOG-level CAT events). Time, latitude, longitude, flight
casting model (ARW-WRF; Skamarock et al. 2008)
level in km, and turbulence intensity of the CAT events
model with the finest grid spacing of 0.37 km is used to
are presented in Table 1, and their horizontal distribution
reproduce this unique cluster of MOG-level CAT events
is depicted in Fig. 1. As shown in Fig. 1, the CAT events
and to understand mechanisms responsible for them in
are distributed in three regions: the west coast, Jeju Island,
the three different areas of South Korea.
and the eastern mountain areas of South Korea. These
The remainder of this paper is organized as follows. In
three areas are indicated by W, J, and E in Table 1,
section 2, observations of nine MOG-level CAT events
respectively. Table 1 shows that most of the CAT events
over South Korea on 2 April 2007 are described. In
occurred in the 3 h from 0530 to 0830 UTC, except for the
section 3, the experimental setting of the WRF model is
first one that occurred at 0200 UTC. CAT events over the
given. In section 4, synoptic features simulated by the
west coast area occurred within 61 h of 0700 UTC. CAT
WRF model and assimilated in the Regional Data As-
events over Jeju Island and the eastern mountain areas
similation and Prediction System (RDAPS) provided by
have more time gaps between individual CAT events.
the Korea Meteorological Administration (KMA) are
Because commercial aircraft usually measure their alti-
described. In section 5, generation mechanisms for the
tudes by barometers, the altitude of each CAT in Table 1
nine MOG-level CAT events that occurred in the three
is calculated using the standard atmospheric assumption
different regions are investigated in detail. Summary
(Lane et al. 2003; Sharman et al. 2006). Among the re-
and conclusions are given in the last section.
ported altitudes, turbulence levels near the west coast
area are relatively lower (z 5 5–8 km) than in the other
2. Investigation of CAT encounters
two regions (z 5 9–13 km). Among the nine CAT events,
On 2 April 2007, 12 turbulence encounters were re- eight are of moderate intensity, whereas the one at
ported in PIREPs over South Korea. Nine of them were 0530 UTC near Jeju Island is a severe-level CAT.
2384 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 2. Composite images of MTSAT focused on (a) Korea and (b) East Asia at 0100 UTC 2 Apr 2007.

To identify the existence of clouds during this pe- 3. Experimental design of WRF model
riod, composite images of the Multifunctional Transport
Satellite (MTSAT) focused on Korea and East Asia are The numerical model used in this study is the ARW-
used (Fig. 2). In Fig. 2a, there are no clouds over the west WRF model, version 3.0, developed by the National
coast and the eastern mountain areas over the Korean Center for Atmospheric Research (NCAR), which is a
Peninsula, while stratiform clouds that extend from South nonhydrostatic and fully compressible model. Details of
China cover Jeju Island. In Fig. 2b, many clouds are lo- this model can be found in Skamarock et al. (2008). This
cated along the surface frontal system elongated from model has been used for several previous studies on
South China to South Japan. Although some patches of turbulence (e.g., Feltz et al. 2009; Fovell et al. 2007; Trier
cloud extend to Jeju Island, there are no signals of pre- and Sharman 2009) and gravity waves (e.g., Plougonven
cipitation in radar echoes or at the surface observation et al. 2008; S.-Y. Kim et al. 2009). Locations of all model
stations over South Korea, including Jeju Island on 2 April domains used in the present simulation are defined by
2007 (not shown). For these reasons, turbulence that oc- the Lambert conformal map projection shown in Fig. 3.
curred in Korea during this period is regarded as CAT, as Six domains (1, 2, 3, 4, 5, and 6) are considered in the
defined by Ellrod et al. (2003). present simulation with horizontal grid spacings of 30,
Before investigating the possible mechanisms of the 10, 3.3, 1.1, 1.1, and 0.37 km, respectively. Note that the
CAT events over South Korea, it should be noted that horizontal grid spacing of domains 4 and 5 is the same.
there are several uncertainties in the PIREP data. First, The locations of domains 1, 2, and 3 and domains 4, 5,
the locations of turbulence in PIREPs sometimes dif- and 6 are depicted in Figs. 3a and 3b, respectively. Also
fer by more than hundreds of kilometers from the actual shown in Fig. 3b are the locations of the operational
locations of the turbulence observed by in situ measure- radiosonde stations over South Korea and the nine
ments (Cornman et al. 2004). Second, the turbulence MOG-level and three light (LGT)-level CAT events
intensity reported in PIREPs tends to be different de- that occurred on 2 April 2007. Two-way nesting in-
pending on the particular pilot and aircraft type or size. teractions are conducted for domains 1 and 2, 2 and 3, 3
The aircraft types of the PIREP data in the present study and 4, and 3 and 5, whereas one-way interaction is con-
are the Boeing 747, Airbus 332, and Airbus 333. Al- ducted for domains 5 and 6. Horizontal grid spacing and
though PIREPs inherently have some uncertainties in the size (5730 3 5130 km) of domain 1 are set exactly the
location and intensity of the turbulence (Schwartz 1996), same as those of the operational RDAPS analysis to al-
they are invaluable resources for validating and evaluat- low direct comparison of synoptic features. Domain 3
ing various CAT diagnostics derived using several oper- includes all CAT events considered in this study, while
ational forecasting models (e.g., McCann 1999; Lee et al. domains 4 and 5 encompass those in Jeju Island and the
2003; Kaplan et al. 2006; Sharman et al. 2006; Knox et al. eastern mountain areas, respectively. Since detailed
2008; Jang et al. 2009; J.-H. Kim et al. 2009). examination of the mountain waves and their breaking is
DECEMBER 2010 KIM AND CHUN 2385

7 km layer to minimize any artificial reflection. Lateral


boundary layers are specified with 5 relaxation grids in
each domain. For initial and boundary conditions, rean-
alysis data from the National Centers for Environmental
Prediction (NCEP) at 6 h intervals and 18 grid spacing in
latitude and longitude (Derber et al. 1991) are used. The
integrations were conducted for 24 h from 1200 UTC 1 to
1200 UTC 2 April 2007 in domains 1 and 2 and for 18 h
from 1800 UTC 1 to 1200 UTC 2 April 2007 in domains
3–6. For efficiency, model time steps are automatically
determined by an adaptive time-stepping method that
selects time steps for all domains to be as large as possi-
ble to satisfy the Courant–Friedrichs–Lewy (CFL) con-
dition.
The parameterizations used in the present simulation

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


include the cloud microphysics scheme by Hong and Lim
(2006), the land surface scheme by Chen and Dudhia
(2001), the Mellor–Yamada–Janjic (MYJ) planetary
boundary layer (PBL) scheme by Janjic (2002), the short
wave radiation scheme by Dudhia (1989), and the long
wave radiation scheme by Mlawer et al. (1997). The cu-
mulus parameterization scheme by Kain (2004) is used
exclusively in domains 1 and 2. Note that the MYJ PBL
scheme takes into account vertical mixing not only in
the PBL but also in the free atmosphere by calculating
subgrid turbulent kinetic energy to third-order accuracy.
Therefore, subgrid-scale turbulence in the CAT regions
can be parameterized by the MYJ PBL scheme without
considering any additional turbulence scheme in the
upper troposphere. To get more realistic model results,
the sixth-order spatial filter (Knievel et al. 2007) is
applied to the model interior, which can damp out 2Dx
waves sufficiently.

4. The synoptic-scale flow


FIG. 3. Horizontal locations of (a) the first three domains (domain 1,
Dx 5 30 km; domain 2, Dx 5 10 km; and domain 3, Dx 5 3.3 km) This section describes the synoptic-scale flow pattern
superimposed on the terrain height (contour interval 5 300 m) of around the Korean Peninsula on 2 April 2007 as it ap-
domain 1 and (b) the last four domains (domain 3; domain 4, Dx 5 peared in the RDAPS analysis data and compares it with
1.1 km; domain 5, Dx 5 1.1 km; and domain 6, Dx 5 0.37 km) su-
the simulated results in the 30-km domain (domain 1).
perimposed on the terrain height (contour interval 5 200 m) of
domain 3. Locations of nine MOG-level and three LGT-level CAT Figure 4 shows horizontal distributions of wind speed
events are additionally depicted in (b) using the conventional with geopotential height at the 300-hPa level (Fig. 4a)
marks of the given turbulence intensities. Operational radiosonde and sea level pressure (Fig. 4b) from the RDAPS analysis
stations over South Korea are depicted in (b) as stars (Baekryungdo, (left) and the WRF simulation results (right) at 0000 UTC
Osan, and Gwangju stations) and dot circles (Sokcho, Pohang,
2 April 2007. At this time, the jet stream flows from
Heuksando, and Gosan stations).
the west to the east over the Korean Peninsula with
a maximum speed of 74 m s21. In addition, a highly
required, an additional finer-resolution (Dx 5 0.37 km, curved trough in the entrance region of the jet stream
where Dx is the horizontal grid spacing) subdomain approaches the Korean Peninsula. As expected from
(domain 6) is included inside of domain 5. the thermal–wind relationship (›U/›z 5 2›T/›y, where
All domains have 87 vertical layers with a vertical grid U and T are the zonal wind and temperature, respec-
spacing of about 300 m and a model top of 20 hPa (z 5 tively), the meridional temperature gradient can be
27–28 km). Rayleigh damping is applied in the uppermost strong over the Korean Peninsula at this time (not
2386 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 4. (a) Geopotential height (gpm; contours) superimposed on horizontal wind speed (m s21; shading) at
300 hPa and (b) sea level pressure (contour) at 0000 UTC 2 Apr 2007 obtained from RDAPS 30-km analysis data
provided by (left) KMA and (right) WRF model output of domain 1. Contour intervals of (a) and (b) are 60 gpm and
4 hPa, respectively.

shown). This is a typical springtime flow pattern in this and Chung 2006; Jang and Chun 2008). The simulated
area, except for the relatively stronger westerly winds results are fairly good compared with the RDAPS analysis
(about 20 m s21 faster) and highly curved trough at in both wind and sea level pressure fields. One noticeable
300 hPa, compared with corresponding fields for the discrepancy, nevertheless, between the analysis and the
6-yr (2003–08) springtime mean calculated by the RDAPS simulated is that the simulated center of the Siberian high
analysis data (not shown). Also, the pressure gradient shifts southward with somewhat different isobaric lines to
force across the Korean Peninsula is exceptionally strong the west of the Siberian high.
on 2 April 2007 because of an intensified Siberian high For further investigation of the synoptic structure and
at the surface. This synoptic feature is similar to that for comparison between the model results and obser-
of CAT events near the upper-level fronts and fronto- vations, vertical profiles of wind and temperature de-
genesis environments shown by previous studies (e.g., rived from the simulations in domains 1–3 and observed
Roach 1970; Keller 1990; Kaplan et al. 2006). The pre- from two radiosonde stations are shown in Fig. 5. Be-
vailing wind is almost perpendicular to the meridionally cause the simulated soundings in the three domains are
oriented mountains located in the eastern part of the similar, only the result in domain 1 will be discussed and
Korean Peninsula, and this provides an optimal envi- compared with the observations. Figure 5 shows the
ronmental condition to generate large-amplitude moun- skew T–logp diagrams of temperature, dewpoint tem-
tain waves (Jiang and Doyle 2004; Doyle et al. 2005; Kim perature, and wind derived using the simulated (gray
DECEMBER 2010 KIM AND CHUN 2387

folding. Details of the tropopause folding will be pre-


sented in section 5a. In addition, the maximum wind
speed of 82 (63.5) m s21 is located at 334 (253) hPa
[about z 5 9 (10.4) km] at Baekryungdo (Gwangju)
station, hence strong VWS below the maximum wind
appears between z 5 4 and 8 km at both stations. The
height range of the enhanced VWS is very similar to that
of the CAT events that occurred over the west coast area
of South Korea on 2 April 2007. The relationship be-
tween the enhanced VWS and the CAT events over the
west coast area will also be examined in section 5a. In
general, the model captures well the aforementioned
vertical structures of the observations, although some
discrepancies exist between the simulation and obser-
vations: (i) low-level wind direction and inversion are

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


not accurately simulated, and (ii) the simulated max-
imum wind speed at z 5 9 km at Baekryungdo is
72.7 m s21, which is smaller than the observed value
of 82 m s21.

5. Model results
a. West coast area
The potential temperatures in domain 2 at z 5 6.1 km
(average level of number 3, 4, and 6 CAT events) at
1800 UTC 1 (left) and 0700 UTC 2 April 2007 (right) are
shown in Fig. 6a, and horizontal wind speed at z 5 9 km
(the level of the maximum jet stream) with horizontal
wind vectors and pressure at z 5 6.1 km are shown in
Fig. 6b. An intensifying upper-level front and associated
jet stream are identified during this 13-h period. At
1800 UTC, a relatively strong meridional temperature
gradient in a frontal zone is located from Balhae Bay
to the northern part of the Korean Peninsula. As this
frontal zone moves southward, the meridional temper-
ature gradient increases at 0700 UTC (Fig. 6b), implying
an intensified upper-level front during this period. As
FIG. 5. Skew T–logp diagrams (solid, temperature; dashed, the strengthening frontal zone induces strong VWS
dewpoint) obtained from the observed (black) and the simulated
(gray) soundings at (a) Baekryungdo and (b) Gwangju observation
through the thermal–wind relationship, the maximum
sites in South Korea at 0000 UTC 2 Apr 2007. (right) Simulated and wind speed of the jet stream above the frontal zone also
(far right) observed wind speeds at each level with full (10 m s21) increases from 70.2 to 82.9 m s21 over the 13 h. In ad-
and half (5 m s21) barbs and a pennant (50 m s21). Isotherms in dition, the convergence zone of north and south west-
108C intervals are denoted by slantwise thin lines in both panels. erlies in the entrance region of the jet stream ahead of
the highly curved trough approaches the west coast of
Korea, where the number 3, 4, and 6 CAT events occurred
lines) and the observed (black lines) soundings at (a) (Fig. 6b). This flow structure is consistent with several
Baekryungdo and (b) Gwangju at 0000 UTC 2 April other CAT events near the upper-level front and the jet
2007. Relatively strong stabilities in the layer between stream (e.g., Keller 1990; Ellrod and Knapp 1992; Kaplan
z 5 4 and 8 km at Baekryungdo and between z 5 3 and et al. 2005).
5 km at Gwangju stations are observed. This intensified To understand how localized turbulence relates to
stability in the midtroposphere is related to the sub- the evolution of the intensifying upper-level front/jet
sidence of stratospheric air because of deep tropopause system, we show in Fig. 7 the horizontal wind speed
2388 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 6. Simulation results in domain 2 at (left) 1800 UTC 1 Apr 2007 and (right) 0700 UTC 2 Apr 2007: (a) potential
temperature (contour) at z 5 6.1 km and (b) horizontal wind vectors and pressure (contour) superimposed on
horizontal wind speed (m s21; shading) at z 5 9 km. Contour intervals of (a) and (b) are 4 K and 4 hPa, respectively.
Locations of CAT events over the west coast area are depicted in the right-hand plots as asterisks with numbers
corresponding to those in Table 1.

(Fig. 7a) and Richardson number (Fig. 7b) with poten- about z 5 10–12 km on the anticyclonic side of the jet
tial temperature and 1.5 potential vorticity unit (PVU) core to about z 5 5 km on the cyclonic side following the
in y–z cross sections along the line shown in Fig. 6b (right), upper-level front. At 0700 UTC (Fig. 7b; right), the
using the same data shown in Fig. 6. At 1800 UTC (Fig. 7a, overall structure of the upper-level front–jet system is
left), the jet core is located at z 5 9 km and the maxi- similar to that at 1800 UTC. However, the slope of the
mum wind speed is 74 m s21. The upper-level front that isentropes near the upper-level front is steeper, so the jet
is represented by slantwise and dense contours of the maximum at z 5 9 km becomes larger (89 m s21). In
potential temperature in the midtroposphere near z 5 addition, the tropopause is deeply folded down to about
3–6 km and y 5 600–1200 km is located below the jet z 5 4 km with a steep 1.5 PVU line following the upper-
core and has relatively strong vertical and horizontal level front. Because of this deep tropopause folding,
gradients of wind speed and temperature, respectively. relatively strong stability layers are located in the mid-
In addition, the height of the dynamic tropopause troposphere because of the intrusion of relatively stable
(1.5 PVU) decreases from its background height of stratospheric air. This is consistent with the observation
DECEMBER 2010 KIM AND CHUN 2389

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 7. Potential temperature (contours) and 1.5 PVU (potential vorticity unit; 1026 K kg21 m2 s21; thick line)
superimposed on (a) horizontal wind speed (m s21; shading) and (b) the Richardson number (shading) along the
dashed line in Fig. 6b (right) at (left) 1800 UTC 1 Apr 2007 and at (right) 0700 UTC 2 Apr 2007. Contour intervals are
4 K. Locations of CAT events over the west coast area are depicted in (b) as asterisks with numbers corresponding to
those in Table 1.

profiles of the temperature shown in Fig. 5. This struc- sections of the Richardson number with the potential
ture is also consistent with tropopause folding that oc- temperature and 1.5 PVU in the different domains are
curred along the upper-level frontal zones in previous presented in Fig. 8. Figures 8a–8c are the simulated results
studies (e.g., Keyser and Shapiro 1986; Lane et al. 2004; from domain 1 (Dx 5 30 km), domain 2 (Dx 5 10 km),
Koch et al. 2005) and is important not only for aviation and domain 3 (Dx 5 3.3 km) in the region occupied by
safety but also for the mixing and stratosphere–troposphere two vertically dashed lines shown in Fig. 7b (right), re-
exchange (STE) of chemical constituents (e.g., Shapiro spectively. Note that the x axis in each plot denotes the
1980; Pavelin et al. 2001). As the development of the relative distance from the southernmost point of each
upper-level front–jet system causes a strong horizon- domain. Overall features of the tropopause folding and
tal temperature gradient in the upper-level frontal zone, upper-level frontal system are similar; except that the
VWS is maximized through the thermal wind rela- lines of 1.5 PVU become steeper as the horizontal grid
tionship. Consequently, the local Richardson number resolution becomes higher. In addition, there is a bump
becomes smaller along this frontal zone, mainly due to at the dynamic tropopause near x 5 700 km and z 5 5–
strong VWS regardless of strong stability (see Fig. 7b). 6 km of Fig. 8c, caused by relatively large-amplitude
The locations and timing of the number 3, 4, and 6 CAT waves generated by steep mountain terrain. The relative
events observed over the west coast of Korea are re- vorticity above the steep mountain near z 5 5–6 km (not
produced in domain 2 of the simulation. shown) is less than 23 3 1024 s21, and consequently,
To investigate the resolution dependency of the sim- potential vorticity has a local minimum there. This is less
ulation results illustrated in Fig. 7b (right), y–z cross evident in low-resolution domains that have smooth and
2390 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 8. Potential temperature (contour), 1.5 PVU (1026 K kg21 m2 s21; thick line), and terrain height (thick line in the bottom) su-
perimposed on the Richardson number (shading) along the region enclosed by two dashed lines at y 5 650 and 1020 km shown in Fig. 7b
(right) in (a) domain 1, (b) domain 2, and (c) domain 3, at 0700 UTC 2 Apr 2007. Contour intervals are 4 K. Note that the x axis in each plot
denotes the relative distance from the southernmost point of each domain. Locations of CAT events over the west coast area are depicted
as asterisks with numbers corresponding to those in Table 1.

small mountains. The regions of the smallest Richardson core, the local Richardson number is less than 0, im-
number along the upper-level front are located below plying that there exists local vertical mixing of potential
the jet core and are mainly due to enhanced VWS along temperature in the left side of Fig. 10c (described below).
the intensifying upper-level front/jet system in all three Figure 9a shows the accumulated convective precipita-
domains. These regions of small Richardson number tion with zero absolute vorticity, horizontal wind vector,
are sustained for about 5 h from 0400 to 0900 UTC, and pressure at the 10.7-km level (averaged level of
during which the intensifying upper-level front/jet system the number 2, 8, and 9 CAT events over the Jeju Island
is dominant over the Korean Peninsula. The smallest area) derived using the simulated results of domain 1 at
Richardson number is 0.224, near the location of the 1900 UTC 1 (left) and 0800 UTC 2 (right). Note that Fig. 9a
number 4 CAT in domain 3. In all three domains, the is not the whole area of domain 1 but focuses on the
regions of the small Richardson number are not obvi- regions of anticyclonic flow and the Jeju Island area.
ously different, implying no significant resolution de- Note also that the absolute vorticity enclosed by the zero
pendency on the simulation results in this area. The lines is negative. During this period, the Jeju Island area
simulation with 30-km horizontal grid spacing can iden- is located on the anticyclonic shear side of the jet streak
tify regions of the CAT events sufficiently in this area at that is passing through the west coast area of South
this time. Korea as described in the previous section. Because of
this strong anticyclonic shear flow, areas of negative
absolute vorticity increase during this period (Fig. 9a).
b. Jeju Island area
Considering that the inertial instability associated with
In Fig. 7b (right), a region of relatively low static sta- the negative absolute vorticity has been regarded as
bility (relatively sparse contours of potential tempera- a possible source of local turbulence that directly affects
ture) and small Richardson number is located on the aircraft in the anticyclonic flow (Knox 1997; McCann
anticyclonic shear side of the jet core near z 5 9–12 km 2001; Trier and Sharman 2009), CAT events observed in
and y 5 300–600 km. The number 2, 8, and 9 CAT events, the Jeju Island area are likely related to the inertial in-
observed over the Jeju Island area, are located in this stability. One can apply Stone (1966)’s criterion for
region at this time in domain 2 (shown in the left of inertial instability—f [ f(1 2 Ri21) 1 z] , 0, where f is
Fig. 10c, described below). This region has relatively the Coriolis parameter, Ri is the Richardson number,
strong horizontal wind shear but weak vertical wind shear and z is the vertical component of relative vorticity.
[vertically oriented contours of isotach; see Fig. 7b (right)]. When we apply Stone’s criterion, the areas of inertial
In some local region on the anticyclonic side of the jet instability are generally similar to those shown in Fig. 9,
DECEMBER 2010 KIM AND CHUN 2391

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 9. Simulated results in domain 1 at (left) 1900 UTC 1 Apr 2007 and at (right) 0800 UTC 2 April 2007: (a) accumulated convective
precipitation (mm; shading) superimposed with horizontal wind barbs, pressure (contours), and zero absolute vorticity (s21; thick lines) at
z 5 10.7 km and (b) NCSU2 index (310212 s23; shading) superimposed with potential temperature (contour) and 0 absolute vorticity
(thick solid line) along the dashed line in (a). Locations of CAT events over the Jeju Island area are depicted in right panels as asterisks
with numbers corresponding to those in Table 1. Contour intervals of (a) and (b) are 4 hPa and 4 K, respectively.

although they occupy somewhat broader regions. This is the accumulated convective precipitation in Fig. 9a (right),
likely because Stone’s criterion includes both symmetric a well-developed convective system appears along the
instability and pure inertial instability. surface frontal zone extending from the mid and southern
As the upper-level front–jet system is intensified dur- parts of China to southern Japan. The overall features of
ing this period, the maximized wind speed of the jet stream the simulated precipitation are in good agreement with
causes the anticyclonic shear region (negative absolute the observed convective bands shown in satellite imagery
vorticity) to be larger near the Jeju Island area (Fig. 9a, (Fig. 2b). As dominant westerly flow passes through the
right). On top of that, ageostrophic flow is also domi- developed convective system, wind vectors are signifi-
nant, which can be seen from the largely different angle cantly perturbed and distorted on the lee side of the
between the wind vectors and isobaric lines. This strong convective system (Fig. 9a, right). As locally enhanced
ageostrophic flow may be caused by well-developed con- cross-streamwise confluent flow (ageostrophic flow)
vective systems in the southern part of China. As seen by forms the local maximum of the vertical component of
2392 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

relative vorticity (vertical vorticity hereafter) on this 2, 8, and 9 CAT events. To understand the linkage be-
anticyclonic shear side of the jet streak region, local tween the developed convective system and inertial in-
concentration of vertical vorticity causes perturbations stability, an additional dry simulation is conducted. The
of absolute vorticity. As a result, the zero contours of experimental setting of the dry simulation is the same as
absolute vorticity are also perturbed (Fig. 9a, right). As- for the control simulation, with two-way nesting inte-
suming that strong horizontal gradients of vertical vor- gration through domains 1–5, except that convective
ticity can be translated to a microscale horizontal vortex parameterization is not used in domains 1 and 2 and
tube (Clark et al. 2000; Kaplan et al. 2005), this could microphysical processes are turned off in all domains.
generate small-scale turbulence that directly affects Figure 10a shows the subgrid-scale TKE with horizontal
aircraft. As can be seen in Fig. 9a (right), the region of wind vectors and boundaries of zero absolute vorticity in
perturbations of zero absolute vorticity along the maxi- the control simulation with moisture process (left) and
mized streamwise-oriented gradient of vertical vorticity dry simulation (right) at 10.7 km at 0800 UTC 2 April
is consistent with the location of the number 2, 8, and 9 2007. Note that the light contours in the left figure are
CAT events over the Jeju Island area. the pressure difference between the control and dry
Figure 9b shows the NCSU2 index with zero absolute simulations. Figures 10b and 10c are the same as Figs. 7a

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


vorticity along the dashed line shown in Fig. 9a. The and 7b, except at 0800 UTC 2 April 2007 in the control
NCSU2 index is maximized where the vectors between (left) and dry (right) simulations. Several interesting
the pressure gradient force and the gradient of stream- features can be found in Fig. 10 by comparing the control
wise vorticity become orthogonal, in contrast to geo- and dry simulations. First, the simulated turbulence
strophic flow where those become parallel. At 1900 UTC (nonzero subgrid-scale TKE) is highly correlated with
1, neither local vertical mixing nor relatively strong val- the inertial instability in the control simulation, while
ues of the NCSU2 index exist in this cross section, even there is no turbulence or inertial instability in the dry
though the negative absolute vorticity region is located in simulation near Jeju Island (Fig. 10a). Second, the jet
the eastern part of China. As the upper-level front–jet streak passing through the central part of the Korean
system intensifies and the dominant westerly flow starts Peninsula is stronger in the control simulation (85 m s21)
to be disturbed as it passes through the well-developed than in the dry simulation (60 m s21) at 10.7-km height
convective system at 0800 UTC 2, local vertical mixing (Fig. 10a). This is mainly because the mesohigh on the
occurs near z 5 9–11 km and x 5 2500–3300 km. This top of the MCS intensifies the meridional pressure gra-
region is well matched with the region of locally maxi- dient on its northern side, and Coriolis acceleration
mized values of the NCSU2 index between z 5 6 and through the geostrophic balance causes the enhanced jet
11 km and x 5 2500 and 3300 km, where the number 2, 8, in the northern or northeastern quadrant of the mesohigh
and 9 CAT events over Jeju Island are located. This (Fig. 10a left). Third, comparing the control simulation
maximized NCSU2 is highly correlated with the negative with the dry simulation, the jet core occurs at a higher
absolute vorticity region, which is induced by ageo- altitude near y 5 700 km and z 5 10.5 km (Fig. 10b), and
strophic flow on the anticyclonic shear side of the jet it causes larger horizontal wind shear on the anticyclonic
streak, as shown in Fig. 9a. Another indicator to identify shear side of the jet core. The minimum Richardson
ageostrophic flow is the Lagrangian Rossby number number associated with inertial instability on the anticy-
[Uccellini et al. (1984); RoL 5 jDy/Dtj/jfyj ’ jy  $yj/jfyj, clonic shear side of the enhanced jet core occurs in the
where y is horizontal wind vector]. The areas of large upper troposphere where the number 2, 8, and 9 CAT
values of the RoL (not shown) are found to coexist in the events are located (Fig. 10c). This is consistent with the
inertial instability along the lee side of the well-developed previous study by Trier and Sharman (2009) who dem-
convective clouds, and they are highly consistent with the onstrated the relationship between the enhanced jet in
areas of large values of NCSU2 index where the number the MCS outflow and aircraft-scale turbulence.
2, 8, 9 CAT events are located over Jeju Island. The re- To identify the resolution dependency of the simu-
gions of relatively large values of NCSU2 and RoL are lated results illustrated in Fig. 10, y–z cross sections of
also consistent with the local maximum of subgrid-scale subgrid-scale TKE superimposed with potential tem-
turbulent kinetic energy (TKE), which will be shown in perature and zero contours of absolute vorticity are
Fig. 10a. As a result, both model-diagnosed turbulence shown in Fig. 11, using the simulated results in domains
potentials and subgrid-scale TKE capture well the regions 1–3 shown in the dashed box in Fig. 10c. Overall features
of CAT events over the Jeju Island area. of the subgrid-scale TKE and potential temperature are
According to Fig. 9, the well-developed convective similar in each domain. However, detailed structures
system played an important role in generating the iner- such as the location of the maximum TKE and areas
tial instability along the ageostrophic flow near the number of negative absolute vorticity are different from each
DECEMBER 2010 KIM AND CHUN 2393

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020

FIG. 10. Comparison between the (left) control and (right) dry simulations at 0800 UTC 2 Apr, 2007: (a) horizontal wind barbs, pressure
(contours on right), pressure difference between the control and dry (contours on left), and zero absolute vorticity (s21; thick contour)
superimposed on the subgrid-scale TKE (m2 s22; shading) in domain 1, (b) potential temperature (contours) and 1.5 PVU (potential
vorticity unit; 1026 K kg21 m2 s21; thick line) superimposed on horizontal wind speed (m s21; shading), and (c) the Richardson number
(shading) along the dashed line in Fig. 6b (right). Contour intervals in (a) are 1 and 4 hPa, respectively. Contour intervals for (b) and (c)
are 4 K. Locations of CAT events over the Jeju Island area are depicted in (a) and (c) as asterisks with numbers corresponding to those in
Table 1.
2394 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 11. Potential temperature (contours) and zero absolute vorticity (thick line; s21 ) superimposed on the subgrid-scale TKE
(m 2 s 22 ; shading) obtained from the simulated results of (a) domain 1, (b) domain 2, and (c) domain 3 focused on the dashed box
in Fig. 10c. Contour intervals are 4 K. Note that the x axis in each plot is the relative distance from the southernmost point of each
domain. Locations of CAT events over the Jeju Island area are depicted as asterisks with numbers corresponding to those in
Table 1.

other, and small-scale features can be seen as horizontal dependent on the structure of topography as well as up-
grid resolution increases. The maximum subgrid-scale stream wind and stability. For this reason, an additional
TKE (1.212 m2 s2) occurs at the location of the number domain, domain 6 (Dx 5 0.37 km), is included in domain
2 CAT in domain 3, whereas it is less close to the number 5 (Dx 5 1.1 km) of the eastern mountain area, as shown in
2 CAT in domain 1. In all domains, the regions of local Fig. 3b, with one-way nesting interaction. Initial condi-
vertical mixing induced by ageostrophically disturbed tions and physical parameterizations in domains 5 and 6
flow on the anticyclonic shear side of the jet core are not are the same as those in the coarser domains, but the
obviously different. This suggests that numerical simu- boundary conditions in domain 6 are from the results of
lation even with 30-km horizontal grid spacing can domain 5 updated every 30 min. Large-scale flow during
identify the regions of CAT events in this area at this this period is in good agreement with observations, al-
time. In addition, the modeled turbulence in this area is ready shown in Figs. 4 and 5.
maintained for about 5 h from 0400 to 0900 UTC in do- Figure 12 shows vertical velocity (shading) with hor-
mains 1, 2, 3, and 4 (not shown). This is because the en- izontal wind vectors in domain 3 at four different levels
vironmental conditions conducive to CAT continue in (z 5 15, 11, 7, and 3 km) at 0300 UTC 2. In this figure,
this area as long as the synoptic or mesoscale forcings the localized wave structure is captured over the eastern
support their generation. These results are somewhat side of South Korea. These waves are generated by the
consistent with those of Trier and Sharman (2009), where strong zonal flow across the meridionally oriented
the simulated lower Richardson number and the subgrid- mountains over the eastern part of the Korean Penin-
scale TKE with 3-km horizontal and 400-m vertical grid sula. In particular, their amplitudes near the locations of
spacings in the anticyclonic outflow of the MCS are sus- the number 1, 5, and 7 CAT events over the eastern
tained for several hours. mountain area are larger than other regions at z 5
11 km, with the maximum amplitude of vertical velocity
of 13.0 and 24.4 m s21 near the location of the number
c. Eastern mountain area
7 CAT. This clearly shows that smaller-scale waves ap-
This section investigates the generation mechanism pear more in the lower altitudes, while only components
of the CAT events that occurred in the eastern moun- with relatively long wavelength reach up to the upper
tain area of South Korea on 2 April 2007. To precisely troposphere.
simulate mountain waves generated by complex terrain, To understand the filtering process of the short-
topography and large-scale flow must be represented re- wavelength waves at low altitude, the vertical profile of
alistically. This is because characteristics of the mountain m2 (where m is the vertical wavenumber) for a station-
waves such as amplitude and wavelength are highly ary mountain wave is calculated using the following
DECEMBER 2010 KIM AND CHUN 2395

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020

FIG. 12. Horizontal wind vectors superimposed on vertical velocity (m s21; shading) obtained from the simulated results of domain 3 at
selected heights z 5 (top left) 15, (top right) 11, (bottom left) 7, and (bottom right) 3 km at 0300 UTC 2 Apr 2007. Locations of CAT
events over the eastern mountain area are depicted as asterisks with numbers corresponding to those in Table 1. Location of domain 6 is
indicated as an inner box in the upper-left plot.

equation: m2 5 (N2/U2) 2 (1/U2)(d2U/dz2) 2 k2, where k 0300 UTC 2 April 2007. The vertical wavenumber of the
is the horizontal wavenumber, and U and N are the basic- stationary gravity waves for a given horizontal wave-
state wind and Brunt–Väisälä frequency, respectively. length changes with altitude as the background stability
Figure 13 shows m2 profiles of waves with horizontal and wind change. Once a vertically propagating gravity
wavelengths of 12.5 km (Fig. 13a) and 4.2 km (Fig. 13b). wave meets a level of m2 5 0, it can be reflected backward
The U and N profiles used to calculate m2 are the domain- from the reflection level (level of m2 5 0) (Gossard and
averaged (domain 6) zonal wind and temperature at Hooke 1975). In Fig. 13, a significant reduction in m2
2396 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 13. Vertical profiles of m2 (where m is the vertical wavenumber of the mountain wave)
with horizontal wavelengths of (a) 12.5 km and (b) 4.2 km calculated using domain-averaged
(domain 6) zonal wind and temperature at 0300 UTC 2 Apr 2007.

appears between z 5 4 and 8 km because of the exis- The 0 wind is derived from the breaking of the gravity
tence of strong variance in background stability in the waves, and it becomes a wave-induced critical level
troposphere, which is mainly due to the deep tropo- (WICL) of the stationary mountain wave (Clark and
pause folding as mentioned in section 5a. There is no Peltier 1984). Once the WICL is formed, consecutive
level of m2 5 0 for the gravity wave with a relatively stationary mountain waves can be either reflected from
longer horizontal wavelength (12.5 km; Fig. 13a), while or absorbed at the critical level, depending on the local
a deep evanescent layer exists (z 5 5.2–10.6 km) for the Richardson number. Since the local Richardson number
wave with short horizontal wavelength (4.2 km; Fig. in the present case is lower than 0.25 near the WICL
13b). For the given U and N used in Fig. 13, we found (Fig. 14a), mountain waves are likely to be reflected.
that waves with horizontal wavelengths shorter than There are two interesting features in Fig. 14d. First,
10.5 km can be reflected in the 4–8-km-height range. almost erect or weakly tilted vertical velocities are shown
Thus, smaller-scale waves appear more in the lower alti- below about z 5 12 km. Spectral analysis of the vertical
tudes, while only relatively long wavelength components velocity with respect to horizontal wavelength reveals
reach up to the upper troposphere, as shown in Fig. 12. that there exists a single spectral peak at 22.1 km for all
Figure 14 shows x–z cross sections of the Richardson heights below z 5 12 km (right side of Fig. 16b, described
number (Fig. 14a) and subgrid-scale TKE with potential below). To examine the possibility of ducting between the
temperature (Fig. 14b), zonal wind (Fig. 14c), and ver- surface and z 5 12 km for a wave with a horizontal wave-
tical velocity (Fig. 14d) in domain 6 at 0300 UTC 2 along length of 22.1 km, the m2 profile of the wave (Fig. 15a) is
the line of y 5 75 km in domain 6 (36.868N in Fig. 12) at calculated using the zonal-mean zonal wind and temper-
which the number 7 CAT is located. Zero and negative ature in the x–z cross section, as shown in Fig. 14d.
values of the wind component are depicted as bold and Figure 15a shows that although m2 decreases with height
dashed lines in Figs. 14c and 14d, respectively. It shows up to about z 5 7 km, it is larger than 0 and increases
that mountain waves simultaneously break down in z 5 above z 5 7 km, with relatively large magnitude at z 5
13.5–17 km. This breaking activates subgrid-scale TKE 12 km. We found that a 0-m2 level exists below z 5 12 km
in that region (Fig. 14b). However, the simulated tur- only for waves with horizontal wavelengths less than
bulence is located above the observed number 7 CAT. 7.7 km. This implies that the wave below z 5 12 km with
This discrepancy will be examined in Fig. 17, which is a 22.1-km horizontal wavelength is not a ducted wave.
described below. Between z 5 3 and 5 (5 and 9) km, Rather, it is a weakly tilted upward-propagating wave.
a relatively strong (weak) stability layer appears, and Figure 15b is the angle of the phase line to the local
this is likely due to the deep tropopause folding, as vertical for the nonhydrostatic
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi internal gravity wave
shown in Fig. 7. In Fig. 14c, the maximum zonal wind, given by cosa 5 k/ k 2 1 m2 , which is calculated using
which exceeds 70 m s21, locates near z 5 8 km. How- the zonal wavenumber corresponding to the 22.1-km
ever, the zonal wind decreases with height above about z horizontal wavelength and the vertical wavenumber ob-
5 10 km, and 0 wind regions appear near z 5 15 km. tained in Fig. 15a. Figure 15b demonstrates clearly that
DECEMBER 2010 KIM AND CHUN 2397

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 14. (a) Richardson number (shading) with potential temperature (contours) and subgrid-scale TKE (turbulent kinetic energy;
m2 s22; shading) with (b) potential temperature, (c) zonal wind (contours), and (d) vertical velocity (contours) obtained from the sim-
ulated result of domain 6 at 0300 UTC 2 Apr 2007 along the line of y 5 75 km in domain 6 at which the No. 7 CAT is located. Contour
intervals in (a) and (b) are 4 K and in (c) and (d) are 10 m s21 and 1 m s21, respectively. Zero and negative values of wind velocity in
(c) and (d) are denoted by thick and dashed contours, respectively. Location of No.7 CAT over the eastern mountain area is depicted as
an asterisk with number corresponding to that in Table 1.

the phase angle is almost 0 below z 5 12 km, while it two regions are the wave-breaking regions shown in
increases above z 5 12 km as m2 increases. The increase Figs. 14a–14c. The waves with horizontal wavelengths
in m2 above z 5 12 km is mainly due to increases in the much smaller than those below z 5 12 km are the sec-
Brunt–Väisälä frequency in the stratosphere. The profiles ondary waves, and there have been several numerical
of m2 and a calculated using the observed sounding at modeling studies on this topic associated with mountain
Osan (not shown) are similar to those shown in Fig. 15. waves (e.g., Bacmeister and Schoeberl 1989; Satomura
This confirms that the waves with horizontal wavelength of and Sato 1999) and convective gravity waves (e.g., Holton
about 22 km below z 5 12 km are vertically propagating and Alexander 1999; Zhou et al. 2002; Snively and Pasko
waves rather than ducted waves. 2003; Lane and Sharman 2006; Chun and Kim 2008).
Second, two regions exist of small horizontal wave- Because of the horizontal resolution, the secondary
lengths in the layer between z 5 12 and 17 km, one near waves appear only in the two high-resolution sub-
x 5 70–100 km and the other near x 5 0–20 km. These domains (domains 5 and 6) with 1.1- and 0.37-km grid
2398 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

FIG. 15. Vertical profiles of (a) m2 (where m is the vertical wavenumber of the mountain

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


wave) and (b) the angle of the wave phase line to the local vertical calculated using the zonal-
mean wind and temperature in the x–z cross section shown in Fig. 14d for a stationary moun-
tain wave with a phorizontal
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi wavelength of 22 km. The angle (a) of the phase line is derived
using cosa 5 k/( k2 1 m2 ), where k and m are the horizontal and vertical wavenumbers,
respectively.

spacings, and waves with horizontal wavelengths less performed an additional simulation with a different
than about 5 km appear only in domain 6. Shown in subgrid-scale turbulence scheme, the Yonsei Univer-
Fig. 16 are the x–z cross sections of vertical velocities sity (YSU) PBL scheme (Noh et al. 2003; Hong et al.
in domains 5 and 6 (left) and their power spectral 2006), in domain 6 through one-way nesting to domain
densities with respect to horizontal wavelength (right). 5 that used the MYJ PBL scheme. The simulated tur-
The vertical velocity in domain 6 is the one shown in bulence near the number 7 CAT using the YSU PBL in
Fig. 14d. In domain 5 (Fig. 16a), the magnitude of the domain 6, however, is not evidently different from that
mountain waves is very small compared with that in using the MYJ PBL scheme (not shown). This implies
domain 6, not only for small-scale secondary waves that that simulated turbulence in domain 6 is not very sensi-
are not fully resolvable but also for the dominant wave tive to the choice of the subgrid-scale turbulence scheme.
with horizontal wavelength of about 22 km below z 5 Rather, simulated variables that are important to de-
12 km. The spectral analysis of the vertical velocity in termine the wave-breaking conditions such as wind and
domain 5 below z 5 12 km reveals that the dominant temperature fields can differ from the observations,
horizontal wavelength is 19.4 km (Fig. 16a right), which and consequently the simulated wave-breaking height
is similar to the dominant horizontal wavelength of is higher than the observed height. To examine this
22.1 km in domain 6 (Fig. 16b right). Although relatively possibility, we estimated the wave-breaking height
small-scale disturbances related to the wave breaking from the mountain-drag parameterization by Chun
appear near x 5 140–160 km and z 5 15–17 km in do- et al. (1996) based on Palmer et al. (1986), using the
main 5, they are much less evident than in domain 6. As basic-state wind and stability profiles obtained from the
the horizontal grid spacing decreases, more complex simulation and observations. The following is a brief
terrain features becomes resolvable and mountain height description from Chun et al. (1996) of how to estimate
generally increases. Consequently, simulated waves have the wave-breaking height: (i) Calculate mountain-wave
a wider spectrum including short wavelength components stress at the surface (denoted by subscript s) t s 5
and larger amplitudes. Since wave breaking is directly 2K0rsNsUsjh9j2, where K0 is the horizontal wavenumber
related to wave amplitude, less or no wave breaking oc- (2.85 3 1024 is used, corresponding to a 22-km horizontal
curs in the relatively low-resolution simulations. wavelength), rs is air density, Ns is the Brunt–Väisälä
Figure 14 clearly demonstrated that regions of wave frequency, Us is the basic-state wind, and jh9j2 is the
breaking exist in the 13.5–17-km-height range, and this variance of subgrid-scale mountain height, which is
can generate local turbulence. However, the simulated calculated in this study using the 30-s digital elevation
wave-breaking regions are located somewhat higher model (DEM) dataset at the nearest grid point of the
than the location of the observed number 7 CAT. To RDAPS analysis data from the number 7 CAT (h9 5
understand the factors that create the discrepancy, we 300 m is used). (ii) Calculate the height variation due
DECEMBER 2010 KIM AND CHUN 2399

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 16. (left) Vertical velocity (m s21) and (right) vertical profiles of log10-scale power spectrum density (PSD) of
the vertical velocity as a function of zonal horizontal wavelength obtained from the simulated results of (a) domain 5
and (b) domain 6 at 0300 UTC 2 Apr 2007. In the left panels, locations of the No. 7 CAT over the eastern mountain
area are depicted with an asterisk.

to the gravity wave at any level using (dh)k 5 [jtk21j/ zonal-mean zonal wind and temperature in the cross
(K0rk21Nk21Uk21)]1/2. (iii) Find the wave-breaking section shown in Fig. 14. The observed basic-state wind
level based on the minimum Richardson number in- and Brunt–Väisälä frequency are derived by applying
cluding wave effects [Rim 5 Ri(1 2 «)/(1 1 «Ri1/2)2], a nine-point vertical average to the rawinsonde data.
where Ri is the local Richardson number, and « is The simulated basic-state wind and stability resembles
the inverse Froude number [« 5 (dh)kNk21/Uk21]. If the observations in a reasonable sense, except with
Rim , 1/ 4, wave breaking occurs and the saturation relatively weak vertical variation and different mag-
wave stress is calculated (Lindzen 1967). If Rim $ 1/ 4, there nitude. Because of this discrepancy, vertical profiles of
is no wave breaking and wave stress is the same as that the minimum Richardson number differ significantly
at the level below (Eliassen and Palm 1960). (iv) Repeat between the simulated and the observed. Based on the
the previous step at higher levels until the wave stress is simulated basic-state flow, the first wave breaking oc-
zero or the model or observation top is reached. curs at z 5 14 km, which is consistent with the altitude
Figure 17 shows vertical profiles of the basic-state wind of the simulated turbulence shown in Fig. 14. In con-
(Fig. 17a), Brunt–Väisälä frequency (Fig. 17b), minimum trast, the wave-breaking height estimated using the
Richardson number (Fig. 17c), and mountain-wave stress observed basic-state flow in the upper troposphere is
(Fig. 17d) derived using the simulated result in domain 12.8 km, which is close to the altitude of the observed
6 (solid) and an observed rawinsonde sounding at number 7 CAT (12.6 km). Detailed inspection of the
Osan (dashed) at 0000 UTC 2 April 2007. Osan is lo- simulated and observed wind and stability near 12–
cated upstream of the eastern mountain area and is 14 km reveals that a slight difference in the vertical
depicted in Fig. 3b. The model-produced basic-state variation and magnitude of the basic-state flow causes
wind and Brunt–Väisälä frequency used in the calcu- a difference in the wave-breaking height of more than
lation of m2 shown in Fig. 15a are obtained from the 1 km.
2400 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


FIG. 17. Vertical profiles of (a) U (m s21), (b) N (s21), (c) minimum Richardson number, and
(d) mountain-wave stress (N m22) obtained from the simulated results of domain 6 (solid) and
the observed soundings at Osan station (dashed) at 0000 UTC 2 Apr 2007. Altitudes of wave
breaking estimated from the simulated (solid lines) and the observed (dashed lines) soundings
are depicted as horizontal lines in (c) and (d).

6. Summary and conclusions the WRF model, including six nested domains with the
finest horizontal grid spacing of 0.37 km. Large-scale
On 2 April 2007, nine MOG-level CAT events were features simulated in the lowest-resolution domain (Dx 5
reported in PIREPs during 6.5 h over the Korean Pen- 30 km) resemble those in the RDAPS analysis data with
insula. This is the highest number of CAT occurrences the same resolution. In addition, vertical profiles for wind
for this region in one day for the last 10 yr. These CAT direction and speed, temperature, and dewpoint tem-
events were distributed in three different regions: the perature derived from the simulations closely match ra-
west coast, Jeju Island, and the eastern mountain areas of diosonde observations.
South Korea. Composite images from the MTSAT satel- Based on the simulation results, we found that gen-
lite reveal no clouds over the west coast and eastern eration mechanisms of the CAT events in the three
mountain areas, whereas the Jeju Island area is covered regions are somewhat different. In the west coast area,
with stratiform clouds extending from the southern part of an intensifying upper-level frontal zone is formed by
China. Based on the available observations such as radar a strong meridional temperature gradient through the
reflectivity and record of precipitation, all CAT events thermal–wind relationship, and upper-level fronto-
including those in Jeju Island occurred in clear or non- genesis causes strengthening of the jet stream over the
convective weather conditions. west coast area, which finally introduces maximized
The characteristics and possible generation mechanisms vertical wind shear below the jet core. Deep tropopause
of the CAT events in the three areas are investigated folding on the cyclonic shear side of the jet core leads to the
using available observations and simulation results of local Richardson number being smaller than 0.25, which
DECEMBER 2010 KIM AND CHUN 2401

can trigger the Kelvin–Helmholtz instability. The lo- can be sufficiently identified by current NWP models
cations of the KHI are consistent with those of the with a horizontal grid spacing of ;10–30 km, while those
observed CAT events over the west coast area. Reso- generated by mountain waves need much higher reso-
lution dependency of the simulated turbulence in this lution. This has an important implication for current and
area is not significant, and a model with 30-km hori- future CAT forecasting. In current CAT forecasting,
zontal grid spacing and 300-m vertical grid spacing can based on the combination of various CAT indices cal-
identify the regions of the CAT events during this period. culated using NWP model output, CAT events associ-
CAT events observed in the Jeju Island area are likely ated with mountain waves cannot be diagnosed well, and
associated with inertial instability. As the intensifying thus alternative approaches may be required. For future
upper-level front/jet system approaches the Korean CAT forecasting, NWP model–forecasted turbulence
Peninsula, the region of inertial instability becomes larger can be used directly instead of calculating diagnostic
and moves to the Jeju Island area in the anticyclonic shear CAT indices, given that turbulence generated by large-
side of the maximized jet core. As strong westerly flows scale forcing is explicitly represented in the numeri-
passing through a well-developed convective system be- cal simulation with a current NWP model resolution.
come disturbed, an ageostrophically induced mesoscale As the resolution of NWP models increases, direct

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


disturbance causes a strong horizontal gradient of local forecasting of CAT from NWP model is likely to be-
vertical vorticity in the lee side of the convective system. come more beneficial.
This results in the local vertical mixings at z 5 9–11 km To advance our understanding of localized turbulence
that directly affects aircraft-scale turbulence in the Jeju in the East Asia region, more CAT cases that occur
Island area. Simulations with 30- and 10-km horizontal under various conditions need to be investigated. In ad-
grid spacings with 300-m vertical grid spacing can detect dition, observational experiments including aircraft in situ
this turbulence with reliable timing and location. measurements over Korea and East Asia may be needed
In the eastern mountain area, CAT events are asso- to understand the mechanisms of CAT generation, like
ciated with breaking of the mountain waves in the lower the previous studies over the Alps (Jiang and Doyle 2004),
stratosphere. As air density decreases with height, the Greenland (Doyle et al. 2005; Ólafsson and Ágústsson
amplitudes of the mountain waves become larger above 2009), and the Pacific Ocean (Lane et al. 2004; Koch et al.
the jet core, and subsequently mountain waves break 2005).
down in a local area and lead to turbulence. Local tur-
Acknowledgments. This work was supported by the
bulence due to subsequent wave breaking over the steep-
Korea Research Foundation Grant funded by the Ko-
est mountain area is identified and reproduced in the
rean Government (MOEHRD, Basic Research Pro-
simulation result with the finest domain (Dx 5 0.37 km),
motion Fund; KRF-2007-313-C00778).
even though the altitude of the simulated turbulence is
somewhat higher than the observation. To determine REFERENCES
the reasons for the altitude discrepancy, the height of the
number 7 CAT over the eastern mountain area is esti- Abarbanel, H. D. I., D. D. Holm, J. E. Marsden, and T. Ratiu, 1984:
Richardson number criterion for the nonlinear stability of
mated using a mountain-wave parameterization with the three-dimensional stratified flow. Phys. Rev. Lett., 52, 2352–2355.
observed and simulated soundings. The results show ——, ——, ——, and ——, 1986: Nonlinear stability analysis of
that a slight difference in the basic-state wind and sta- stratified fluid equilibria. Philos. Trans. Roy. Soc. London,
bility between the simulation and observations near the A318, 349–409.
Bacmeister, J. T., and M. R. Schoeberl, 1989: Breaking of vertically
CAT cause a significant discrepancy in the altitudes of
propagating two-dimensional gravity waves forced by orog-
the turbulence (more than 1 km). raphy. J. Atmos. Sci., 46, 2109–2134.
Although several discrepancies exist in the location Chen, F., and J. Dudhia, 2001: Coupling an advanced land surface–
and timing of individual CAT events between simula- hydrology model with the Penn State–NCAR MM5 modeling
tion and observations, the generation mechanisms of the system. Part I: Model implementation and sensitivity. Mon.
observed CAT events in different regions are properly Wea. Rev., 129, 569–585.
Chun, H.-Y., and Y.-H. Kim, 2008: Secondary waves generated by
reproduced in the present simulations using the WRF breaking of convective gravity waves in the mesosphere and
model. It is noteworthy, however, that the capability of their influence in the wave momentum flux. J. Geophys. Res.,
numerical simulation of individual CAT depends not 113, D23107, doi:10.1029/2008JD009792.
only on the model resolution but also on its generation ——, J.-H. Jung, J.-H. Oh, and J.-W. Kim, 1996: Effects of mountain-
mechanism. CAT events that are generated by large- induced gravity wave drag on atmospheric general circulation.
J. Korean Meteor. Soc., 32, 581–591.
scale forcings such as upper-level frontogenesis and in- Clark, T. L., and W. R. Peltier, 1984: Critical level reflection and
ertial instability associated with ageostrophic flow, as the resonant growth of nonlinear mountain waves. J. Atmos.
shown in the present west coast and Jeju Island areas, Sci., 41, 3122–3134.
2402 JOURNAL OF APPLIED METEOROLOGY AND CLIMATOLOGY VOLUME 49

——, W. D. Hall, R. M. Kerr, D. Middleton, L. Radke, F. M. Ralph, Turbulence Forecasting Algorithm (KITFA). Atmosphere,
P. J. Nieman, and D. Levinson, 2000: Origins of aircraft 19, 255–268.
damaging clear-air turbulence during the 9 December 1992 Janjic, Z. I., 2002: Nonsingular implementation of the Mellor-Yamada
Colorado downslope windstorm. J. Atmos. Sci., 57, 1105–1131. level 2.5 scheme in the NCEP Meso Model. NCEP Office Note
Cornman, L. B., G. Meymaris, and M. Limber, 2004: An update 437, 61 pp.
on the FAA Aviation Weather Research Program’s in situ Jiang, Q., and J. D. Doyle, 2004: Gravity wave breaking over the
turbulence measurement and reporting system. Preprints, central Alps: Role of complex terrain. J. Atmos. Sci., 61, 2249–
11th Conf. on Aviation, Range, and Aerospace Meteorology, 2266.
Hyannis, MA, Amer. Meteor. Soc., P4.3. [Available online at Joseph, B., A. Mahalov, B. Nicolaenko, and K. L. Tse, 2004: Var-
http://ams.confex.com/ams/11aram22sls/techprogram/paper_ iability of turbulence and its outer scales in a model tropo-
81622.htm.] pause jet. J. Atmos. Sci., 61, 621–643.
Derber, J. C., D. F. Parrish, and S. J. Lord, 1991: The global Kain, J. S., 2004: The Kain–Fritsch convective parameterization:
operational analysis system at the National Meteorological An update. J. Appl. Meteor., 43, 170–181.
Center. Wea. Forecasting, 6, 538–547. Kaplan, M. L., A. W. Huffman, K. M. Lux, J. D. Cetola, J. J.
Doyle, J. D., M. A. Shapiro, Q. Jiang, and D. L. Bartels, 2005: Charney, A. J. Riordan, Y.-L. Lin, and K. T. Waight III,
Large-amplitude mountain wave breaking over Greenland. 2005: Characterizing the severe turbulence environments as-
J. Atmos. Sci., 62, 3106–3126. sociated with commercial aviation accidents. Part II: Hydro-
Dudhia, J., 1989: Numerical study of convection observed during static mesoscale numerical simulations of supergradient wind

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


the Winter Monsoon Experiment using a mesoscale two- flow and streamwise ageostrophic frontogenesis. Meteor. Atmos.
dimensional model. J. Atmos. Sci., 46, 3077–3107. Phys., 88, 129–153.
Eliassen, A., and E. Palm, 1960: On the transfer of energy in sta- ——, and Coauthors, 2006: Characterizing the severe turbulence
tionary mountain waves. Geophys. Publ., 22, 1–23. environments associated with commercial aviation accidents.
Ellrod, G. P., and D. I. Knapp, 1992: An objective clear-air tur- A real-time turbulence model (RTTM) designed for the op-
bulence forecasting technique: Verification and operational erational prediction of hazardous aviation turbulence envi-
use. Wea. Forecasting, 7, 150–165. ronments. Meteor. Atmos. Phys., 94, 235–270.
——, and J. A. Knox, 2010: Improvements to an operational clear- Keller, J. L., 1990: Clear-air turbulence as a response to meso- and
air turbulence diagnostic index by addition of a divergence synoptic-scale dynamic processes. Mon. Wea. Rev., 118, 2228–
trend term. Wea. Forecasting, 25, 789–798. 2242.
——, P. F. Lester, and L. J. Ehernberger, 2003: Clear-air turbu- Keyser, D., and M. A. Shapiro, 1986: A review of the structure
lence. Encyclopedia of the Atmospheric Sciences, J. R. Holton and dynamics of upper-level frontal zones. Mon. Wea. Rev.,
et al., Eds., Academic Press, 393–403. 114, 452–499.
Feltz, W. F., K. M. Bedka, J. A. Otkin, T. Greenward, and Kim, J.-H., and I.-U. Chung, 2006: Study on mechanisms and
S. A. Ackerman, 2009: Understanding satellite-observed orographic effect for the springtime downslope windstorm
mountain-wave signatures using high-resolution numerical over the Yeongdong region. Atmosphere, 16, 67–83.
model data. Wea. Forecasting, 24, 76–86. ——, and H.-Y. Chun, 2008: Analysis on the spatial and temporal
Fovell, R. G., R. D. Sharman, and S. B. Trier, 2007: A case study distribution of the aircraft turbulence occurred in South Korea
of convectively-induced clear-air turbulence. Preprints, 12th for the recent 10 years. Preprints, 13th Conf. on Aviation,
Conf. on Mesoscale Processes, Waterville Valley, NH, Amer. Range, and Aerospace Meteorology, New Orleans, LA, Amer.
Meteor. Soc., 13.4. [Available online at http://ams.confex.com/ Meteor. Soc., P3.3. [Available online at http://ams.confex.
ams/pdfpapers/126190.pdf.] com/ams/88Annual/techprogram/paper_128033.htm.]
Gossard, E. E., and W. H. Hooke, 1975: Waves in the Atmosphere. ——, and ——, 2011: Statistics and possible sources of aviation tur-
Elsevier, 456 pp. bulence over South Korea. J. Appl. Meteor. Climatol., in press.
Hines, C. O., 1960: Internal atmospheric gravity waves at iono- ——, ——, W. Jang, and R. D. Sharman, 2009: A study of forecast
spheric heights. Can. J. Phys., 38, 1441–1481. system for clear-air turbulence in Korea. Part II: Graphical
Holton, J. R., and M. J. Alexander, 1999: Gravity waves in the Turbulence Guidance (GTG) system. Atmosphere, 19, 269–287.
mesosphere generated by tropospheric convection. Tellus, Kim, S.-Y., H.-Y. Chun, and D. L. Wu, 2009: A study on strato-
51B, 45–58. spheric gravity waves generated by Typhoon Ewiniar: Nu-
Hong, S.-Y., and J.-O. J. Lim, 2006: The WRF single-moment 6-class merical simulations and satellite observations. J. Geophys.
microphysics scheme (WSM6). J. Korean Meteor. Soc., 42, Res., 114, D22104, doi:10.1029/2009JD011971.
129–151. Knievel, J. C., G. H. Bryan, and J. P. Hacker, 2007: Explicit nu-
——, Y. Noh, and J. Dudhia, 2006: A new vertical diffusion merical diffusion in the WRF model. Mon. Wea. Rev., 135,
package with an explicit treatment of entrainment processes. 3808–3824.
Mon. Wea. Rev., 134, 2318–2341. Knox, J. A., 1997: Possible mechanism of clear-air turbulence
Howard, L. N., 1961: Note on a paper by John Miles. J. Fluid Mech., in strongly anticyclonic flow. Mon. Wea. Rev., 125, 1251–
10, 509–512. 1259.
Jaeger, E. B., and M. Sprenger, 2007: A Northern Hemispheric ——, D. W. McCann, and P. D. Williams, 2008: Application of the
climatology of indices for clear-air turbulence in the tropo- Lighthill–Ford theory of spontaneous imbalance to clear-air
pause region derived from ERA-40 reanalysis data. J. Geo- turbulence forecasting. J. Atmos. Sci., 65, 3292–3304.
phys. Res., 112, D20106, doi:10.1029/2006JD008189. Koch, P., H. Wernli, and H. W. Davies, 2006: An event-based
Jang, W., and H.-Y. Chun, 2008: Severe downslope windstorms jet-stream climatology and typology. Int. J. Climatol., 26,
of Gangneung in the springtime. Atmosphere, 18, 207–224. 283–301.
——, ——, and J.-H. Kim, 2009: A study of forecast system Koch, S. E., and Coauthors, 2005: Turbulence and gravity waves
for clear-air turbulence in Korea. Part I: Korean Integrated within an upper-level front. J. Atmos. Sci., 62, 3885–3908.
DECEMBER 2010 KIM AND CHUN 2403

Lane, T. P., and R. D. Sharman, 2006: Gravity wave breaking, Prusa, J. M., P. K. Smolarkiewicz, and R. R. Garcia, 1996: Propa-
secondary wave generation, and mixing above deep convec- gation and breaking at high altitudes of gravity waves excited
tion in a three-dimensional cloud model. Geophys. Res. Lett., by tropospheric forcing. J. Atmos. Sci., 53, 2186–2216.
33, L23813, doi:10.1029/2006GL027988. Roach, W. T., 1970: On the influence of synoptic development on
——, and ——, 2008: Some influences of background flow condi- the production of high level turbulence. Quart. J. Roy. Meteor.
tions on the generation of turbulence due to gravity wave Soc., 96, 413–429.
breaking above deep convection. J. Appl. Meteor. Climatol., Satomura, T., and K. Sato, 1999: Secondary generation of gravity
47, 2777–2796. waves associated with the breaking of mountain waves. J. Atmos.
——, ——, T. L. Clark, and H.-M. Hsu, 2003: An investigation Sci., 56, 3847–3858.
of turbulence generation mechanisms above deep convection. Schwartz, B., 1996: The quantitative use of PIREPs in developing avi-
J. Atmos. Sci., 60, 1297–1321. ation weather guidance products. Wea. Forecasting, 11, 372–384.
——, J. D. Doyle, R. Plougonven, M. A. Shapiro, and R. D. Sharman, Sekioka, M., 1970: Application of Kelvin–Helmholtz instability to
2004: Observations and numerical simulations of inertia–gravity clear-air turbulence. J. Appl. Meteor., 9, 896–899.
waves and shearing instabilities in the vicinity of a jet stream. Shapiro, M. A., 1980: Turbulent mixing within tropopause folds
J. Atmos. Sci., 61, 2692–2706.
as a mechanism for the exchange of chemical constituents
——, ——, R. D. Sharman, M. A. Shapiro, and C. D. Watson, 2009:
between the stratosphere and troposphere. J. Atmos. Sci., 37,
Statistics and dynamics of aircraft encounters of turbulence
994–1004.
over Greenland. Mon. Wea. Rev., 137, 2687–2702.

Downloaded from http://journals.ametsoc.org/doi/pdf/10.1175/2010JAMC2449.1 by guest on 17 June 2020


Sharman, R., C. Tebaldi, G. Wiener, and J. Wolff, 2006: An in-
Lee, Y.-G., B.-C. Choi, R. Sharman, G. Wiener, and H.-W. Lee,
tegrated approach to mid- and upper-level turbulence fore-
2003: Determination of the primary diagnostics for the CAT
(clear-air turbulence) forecast in Korea. J. Korean Meteor. casting. Wea. Forecasting, 21, 268–287.
Soc., 39, 677–688. Skamarock, W. C., J. B. Klemp, J. Dudhia, D. O. Gill, D. M. Barker,
Lilly, D. K., 1978: A severe downslope windstorm and aircraft M. G. Duda, X.-Y. Huang, W. Wang, and J. G. Powers, 2008: A
turbulence event induced by a mountain wave. J. Atmos. Sci., description of the Advanced Research WRF version 3. NCAR
35, 59–77. Tech. Note NCAR/TN-4751STR, 88 pp.
Lindzen, R. S., 1967: Thermally driven diurnal tide in the atmo- Smith, R. B., 1979: The influence of mountains on the atmosphere.
sphere. Quart. J. Roy. Meteor. Soc., 93, 18–42. Advances in Geophysics, Vol. 21, Academic Press, 87–230.
Mancuso, R. L., and R. M. Endlich, 1966: Clear-air turbulence ——, 1989: Hydrostatic air-flow over mountains. Advances in
frequency as a function of wind shear and deformation. Mon. Geophysics, Vol. 31, Academic Press, 1–41.
Wea. Rev., 94, 581–585. ——, S. Skubis, J. D. Doyle, A. S. Broad, C. Kiemle, and
McCann, D. W., 1999: A simple turbulent kinetic energy equa- H. Volkert, 2002: Mountain waves over Mont Blanc: In-
tion and aircraft boundary layer turbulence. Natl. Wea. Dig., fluence of a stagnant boundary layer. J. Atmos. Sci., 59,
23, 13–19. 2073–2092.
——, 2001: Gravity waves, unbalanced flow, and aircraft clear-air Snively, J. B., and V. P. Pasko, 2003: Breaking of thunderstorm-
turbulence. Natl. Wea. Dig., 25, 3–14. generated gravity waves as a source of short-period ducted
Miles, J. W., 1961: On the stability of heteorogeneous shear flows. waves at mesopause altitudes. Geophys. Res. Lett., 30, 2254,
J. Fluid Mech., 10, 496–508. doi:10.1029/2003GL018436.
——, 1986: Richardson’s criterion for the stability of stratified Stone, P. H., 1966: On nongeostrophic baroclinic stability. J. Atmos.
shear flow. Phys. Fluids, 29, 3470–3471. Sci., 23, 390–400.
Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and Trier, S. B., and R. D. Sharman, 2009: Convection-permitting
S. A. Clough, 1997: Radiative transfer for inhomogeneous simulations of the environment supporting widespread tur-
atmospheres: RRTM, a validated correlated-k model for the bulence within the upper-level outflow of a mesoscale con-
longwave. J. Geophys. Res., 102, 16 663–16 682. vection system. Mon. Wea. Rev., 137, 1972–1990.
Noh, Y., W. G. Cheon, S.-Y. Hong, and S. Raasch, 2003: Im- ——, ——, R. G. Fovell, and R. G. Frehlich, 2010: Numerical
provement of the K-profile model for the planetary boundary
simulation of radial cloud bands within the upper-level out-
layer based on large eddy simulation data. Bound.-Layer
flow of an observed mesoscale convective system. J. Atmos.
Meteor., 107, 401–427.
Sci., 67, 2990–2999.
Ólafsson, H., and H. Ágústsson, 2009: Gravity wave breaking in
Uccellini, L. W., P. J. Kocin, R. A. Petersen, C. H. Wash, and
easterly flow over Greenland and associated low level barrier-
K. F. Brill, 1984: The Presidents’ Day cyclone of 18–19 February
and reverse tip-jets. Meteor. Atmos. Phys., 104, 191–197.
Palmer, T. N., G. J. Shutts, and R. Swinbank, 1986: Alleviation of a 1979: Synoptic overview and analysis of the subtropical jet
systematic westerly bias in general circulation and numerical streak influencing the pre-cyclogenetic period. Mon. Wea. Rev.,
weather prediction models through an orographic gravity 112, 31–54.
wave drag parameterization. Quart. J. Roy. Meteor. Soc., 112, VanZandt, T. E., and D. C. Fritts, 1989: A theory of enhanced
1001–1039. saturation of the gravity wave spectrum due to increases in
Pavelin, E., J. A. Whiteway, and G. Vaughan, 2001: Observations atmospheric stability. Pure Appl. Geophys., 130, 399–420.
of gravity wave generation and breaking in the lowermost Wurtele, M. G., R. D. Sharman, and A. Datta, 1996: Atmospheric
stratosphere. J. Geophys. Res., 106, 5173–5179. lee waves. Annu. Rev. Fluid Mech., 28, 429–476.
Plougonven, R., A. Hertzog, and H. Teitelbaum, 2008: Observa- Zhou, X., J. R. Holton, and G. L. Mullendore, 2002: Forcing
tions and simulations of a large-amplitude mountain wave of secondary waves by breaking of gravity waves in the
breaking over the Antarctic Peninsula. J. Geophys. Res., 113, mesosphere. J. Geophys. Res., 107, 4058, doi:10.1029/
D16113, doi:10.1029/2007JD009739. 2001JD001204.

You might also like