Charney and Keyser - Double Trouble_20071107

advertisement
Mesoscale model simulations of the meteorological conditions during the
2 June 2002 Double Trouble State Park wildfire
Joseph J. Charney and Daniel Keyser
Introduction
As a thick morning mist cleared on 2 June 2002, the New Jersey Forest Fire
Service (NJFFS) lookout at the Cedar Bridge Fire Tower in east-central New Jersey (NJ)
reported a smoke plume in the northeastern corner of the Double Trouble State Park
(DT), near the Jake’s Branch of the Toms River (Figs. 1a,b) at 1709 UTC (NJFFS 2003).
Firefighters from two NJFFS units responded immediately to the fire, arriving on the
scene within five minutes of the initial report, at 1714 UTC (Table 1). Eleven minutes
later, at 1725 UTC, the firefighters were forced to abandon their initial attack and pull
back to establish defensible positions around the town of Beachwood, NJ, less than 3 km
downwind of the fire. By the sunset that evening, 361 wildland firefighters had been
called in to fight a fire that burned 1,300 acres (Fig. 2), forced the closure of the Garden
State Parkway, damaged or destroyed 36 homes and outbuildings, directly threatened
over 200 homes, forced the evacuation of 500 homes, and caused an estimated $400,000
in property damage (NJFFS 2003).
Prior to the initial attack, at 1415 UTC, weather observations at the Lakewood
Fire Tower indicated surface winds of 4 ms-1, temperatures near 24°C, and a relative
humidity of 62%. Approximately two hours after the initial attack (1900 UTC), a spot
weather forecast prepared by the Mount Holly, NJ, National Weather Service (NWS)
1
Forecast Office (KMHO) predicted winds of 7 ms-1 with gusts to 13 ms-1, a temperature
of 28°C , and a relative humidity of 28% (Table. 2). The Lakewood Fire Tower reported
winds in excess of 18 ms-1 at 1725 UTC, coinciding with the time when the firefighters
were forced to abandon their initial attack on the fire. These observations indicate that a
profound change occurred in the atmosphere between the late morning and the early
afternoon of 2 June, resulting in a substantial drop in relative humidity accompanied by
an increase in surface wind speeds at the fire location.
The Eastern Area Modeling Consortium (EAMC), in East Lansing, MI, operates a
real-time mesoscale modeling system that predicts the weather conditions throughout the
north-central and northeastern United States (Heilman et al. 2003; Zhong et al. 2005). At
the time of the fire, EAMC simulations indicate that dry, high-momentum air at the
surface was most evident over southeastern Pennsylvania (PA) and southern NJ. Since
the horizontal distribution of moisture cannot be attributed to a land-sea boundary or
other local land surface characteristics, the generation mechanism for the moisture and
wind anomalies are hypothesized to derive from larger than local scale (e.g. mesoscale)
atmospheric conditions at the ground and aloft.
Kaplan et al. (2008) analyze the dynamics of the mesoscale circulations
associated with the generation of dry air aloft during the DT fire, concentrating on the
physical processes that contribute to the development of the observed surface weather
conditions. Our paper focuses on the diagnostic capabilities of the observations and
numerical weather prediction models available in real-time at the time of the fire. We
highlight connections between atmospheric conditions aloft and the evolution of the
surface conditions that may have impacted the observed fire behavior. We do not attempt
2
to simulate the evolution of the fire. Instead, we focus on the utility of the available
observations and the ability of a mesoscale model to predict and diagnose the local
weather conditions that contributed to the observed fire behavior during the DT fire. In
the next section we describe the mesoscale atmospheric model employed by the EAMC
to study this event. Following sections investigate the observed fire behavior and
meteorological conditions, and present mesoscale model simulations that diagnose
weather conditions that contributed to the observed fire behavior. The final section
consists of discussion and conclusions.
Mesoscale model description
The EAMC mesoscale modeling system is built around the Penn State
University/National Center for Atmospheric Research Mesoscale Model version 5.3
(MM5) (Grell et al. 1995). The MM5 is a nonhydrostatic, primitive-equation, mesoscale
model that uses a terrain-following pressure coordinate system. The EAMC formulation
of the model employs physical subgrid-scale parameterizations to account for radiative
transfer (Mlawer et al. 1997), mixed-phase cloud microphysics (Reisner et al. 1998),
cumulus convection (Kain and Fritsch 1990), boundary-layer turbulence (Janjic 1990),
and land-surface exchange processes (Chen and Dudhia, 2001). The EAMC runs the
MM5 twice daily initialized at 0000 and 1200 UTC using the National Centers for
Environmental Prediction (NCEP) operational Eta model output. The model produces
48-h simulations over 36- and 12-km domains, and 24-h simulations for two 4-km
domains. The model output is available hourly for display and analysis purposes and the
3
hourly output is archived for forecast validation. The outermost 36-km domain covers
the entire continental United States, with a one-way nested 12-km domain covering the
New England, the Mid-Atlantic states, the Ohio Valley, and the Great Lakes region. Two
additional one-way nested 4-km domains are generated from the 12-km simulation, one
covering the Great Lakes Region and another covering most of New England and the
Mid-Atlantic states (see Fig. 1 of Zhong et al. 2005). The model employs 35 vertical
levels, with 15 vertical levels located in the lowest 2000 m to better resolve planetary
boundary layer (PBL) structures that have the most direct impact on fire behavior.
Although the real-time simulation using Eta initial conditions reproduces many of
the observed surface phenomena associated with the DT fire, the simulation fails to
adequately represent the PBL depths observed at the time of the fire. Since PBL
processes are important to fully understanding the impact of the overlying atmosphere on
the observer fire behavior, we performed additional simulations initialized using the
NCEP North American Regional Reanalysis (NARR) (Mesinger et al. 2006). While
these simulations employ data that was not available in real-time during the fire incident,
the results of the simulation and ensuing analysis can be used to inform real-time
simulation results and thus impact decision-making during a wildfire.
Observed fire behavior and meteorological conditions
The DT wildfire started from a camp fire that had been left smoldering on the
night of 1 June and was small enough to escape detection at fire observation towers until
1709 UTC (Table 1) due to poor visibility and high relative humidity in the morning
4
(Steve Maurer, NJFFS, personal communication). The Keech-Byram Drought Index
(Keech and Byram 1968) in southern NJ at the time of the fire was 105, which indicates a
long-term, climatological drought at the “moderate” level. Other indices suggest that the
fuel conditions were at moderate risk for fire. However, NJFFS (2003) reports that frost
damage was present in the area due to unusual freezing temperatures that occurred some
weeks before the fire. As a result, the drought and fuel indices for the area may have
underestimated the amount of dead and dry fuel that existed at the time and location of
the DT fire.
Table 1 summarizes the observed weather conditions, fire behavior, and
firefighting activities reported by the NJFFS throughout the day on 2 June (NJFFS 2003).
Within 16 minutes of the initial observation and 11 minutes after the initial attack, the fire
was found to be growing too rapidly to be contained. Wind speeds at the Lakewood Fire
Tower were reported to be in excess of 18 ms-1 at 1725 UTC. Radio logs from the
Lakewood Fire Tower indicate that the fire was advancing at >34 m per minute during
the initial attack. The sudden change in surface weather conditions and the resulting
extreme fire behavior forced the fire crews to retreat. The crews lit back fires ahead of
the fire front in an effort to deprive the fire of fuel and prevent it from sweeping through
the town of Beachwood less than 3 km to the east. A combination of a surface wind shift
from westerly to northwesterly and the actions of the firefighters prevented the fire from
entering the town. Continued efforts were required throughout the day to protect homes,
outbuildings, and other structures to the south of the town as the fire continued to spread
rapidly, with multiple surface wind shifts and spot fires occurring ahead of the fire front.
5
The weather observations available at the time of the fire give ample evidence of
variable atmospheric conditions throughout the day. A synoptic analysis of the observed
surface winds reveals that winds in eastern PA and NJ are generally light, with westerly
and west-southwesterly winds dominating at 1200 UTC (Fig. 3a). Further to the north
and west, a wind shift associated with a dry front produces northwesterly or northnorthwesterly flow. At 1800 UTC (Fig. 2b), which coincides with the time when the fire
was declared a “Major Fire” and within an hour of wind shifts starting to occur on the fire
line (Table 1), the surface analysis indicates the dry front passing through south-central
NJ, with westerly winds predominating in NJ. Also note that the wind speeds in
Maryland (MD), NJ, and coastal New England had increased by this time, from 10 kts (5
ms-1) to 25 kts (13 ms-1) in southern NJ. By 0000 UTC (Fig. 3c), winds across MD, PA,
and NJ were generally northwesterly or north-northwesterly, with speeds around 5 kts (3
ms-1). Surface mixing ratios in southern NJ fall from 12 g/kg at 1200 UTC to5 g/kg at
1800 UTC and 7 g/kg at 0000 UTC (Fig. 2a-c). While a drying trend is consistent with
the frontal passage evident in the wind field, note that isolated areas of low mixing ratios
(5 g/kg) develop in southern NJ as well as in south-central PA and northern MD at 1800
UTC. The spot forecast issued by KMHO for the fire location shows the same trend in
wind speed and moisture (relative humidity) through the afternoon of 2 June, and
indicates that wind speeds were expected to remain high through the evening as the
relative humidity increased, and then taper off after sunset.
The conditions aloft in the northeastern United States broadly indicate the
influence of an upper level trough moving through New England. At 1200 UTC, strong
northerly flow in New England and a sharp temperature gradient along the St. Lawrence
6
River is evident at 850 hPa (Fig. 4a). At 300 hPa, a jet streak north and west of the Great
Lakes moves downstream into southern New England and the Mid-Atlantic states (Fig.
4b). By 0000 UTC, the sharp temperature gradient along the St. Lawrence River has
decreased as the trough progresses southward and slightly eastward. Nevertheless,
northerly flow and cooler temperatures are evident at 850 hPa in New York (NY),
northern PA, and northern NJ (Fig. 4c). At 300 hPa, the jet streak has advanced to the
east and south over southern New England while the magnitude of the wind maximum
has decreased in the last 6 hours (Fig. 4d). These upper level analyses support the surface
analyses indicating the presence of a weak, dry cold front moving through southern New
England and the Mid-Atlantic states during the afternoon and early evening of 2 June. .
Geostationary Operational Environmental Satellites (GOES) water vapor imagery
for 1215 UTC 2 June and 0015 UTC 3 June (Figs. 5a,b) delineates a ribbon of dry air
sagging southward and eastward during this time period, coincident with the jet streak
depicted in Figs. 4b,d. The GOES observations suggest that a ribbon of dry air is present
in the mid- to upper-troposphere at this time, and that the prevailing synoptic and
mesoscale flow patterns (i.e. the upper level trough and jet streak) propagate this dry air
through the area of interest at the time of the fire. A MODIS visible satellite image valid
at 1558 – 1611 UTC 2 June supports the assertion that dry air was present over New
Jersey near the time when the extreme fire behavior was reported (Fig. 5c).
A skew-T/log-P sounding from 1200 UTC 2 June at Upton, NY (UKX) (Fig. 6a)
indicates a shallow surface-based mixed layer beneath a thermal inversion above 950
hPa. Near-surface winds are light and from the southwest. Above a inversion, a deeper
well-mixed layer extends to 550 hPA with 35 kt (18 ms-1) winds from the west-
7
southwest prevailing throughout the layer. Above 550 hPa, the air is considerably dryer
and wind speeds increase to 75 kts (39 ms-1). The 0000 UTC 2 June sounding at UKX
(Fig. 6b) reveals that while the surface temperature is close to what was observed 6 h
earlier, the dewpoint is lower by 10°C. The surface-based mixed layer has deepened to
about 775 hPa, and the dry air aloft is just above the inversion at 750 hPa. Winds
throughout the profile are now from the north-northwest, with surface wind speeds of 10
kts (5 ms-1) increasing gradually to 60 kts (31 ms-1) at 600 hPa. This change is consistent
with the surface and upper air observations, in that surface drying occurs coincident with
an increase in surface wind speed and a wind shift from southwesterly to northwesterly.
The soundings also suggest that dry, high-momentum air subsides throughout the day,
such that a considerable contrast develops between the mixed-layer air near the ground
and the air above the inversion during the 6-hour period when the fire occurred.
Wind profiler observations from New Brunswick, NJ (Fig. 7) show the evolution
of the near-surface winds at the intervening times. Note that between 1400 UTC and
1900 UTC, wind speeds in the lowest 500 m increase from 5 ms-1 to 15 ms-1, coincident
with the time when the fire started to grow rapidly. Surface meteograms from Atlantic
City, NJ (ACY) (Fig. 8a) and McGuire Air Force Base, NJ (WRI) (Fig. 8b), also show
that surface wind speeds and gustiness increase between 1500 and 1800 UTC, although
ACY also reports increasing wind speeds and gustiness 2 h earlier. Both stations report a
dewpoint minimum at 1800 UTC, which coincided with the time when the NJFFS
declared this incident to be a “major fire” (Table 1). The wind shift and gradual decrease
in dew point temperature between 1300 and 1800 UTC are consistent with the passage of
a dry front.
8
The frontal passage diagnosed in the surface observations, the dry ribbon aloft
evident in satellite imagery, and the evolution of the upper level trough could be critical
pieces of evidence for explaining the surface drying and increase in surface wind speeds
and gustiness observed at the time of the outbreak and spread of the fire. It is
hypothesized that coupling between a deepening mixed layer and a midtropospheric dry
layer occurs in the post-frontal environment and leads to the downward mixing of dry air
and momentum in the PBL. Since strong winds (Byram 1954; Fahnstock 1965; Brotak
and Reifsnyder 1977; Simard et al. 1987) and low relative humidity (Lansing 1939; Davis
1969; Simard et al. 1987) are known to be conducive to large fire development and rapid
fire spread, this sequence of meteorological events could help explain the observed fire
behavior in this case.
Mesoscale model simulations
While the observed weather conditions provide evidence that unusually dry and
windy conditions developed in southern NJ on the afternoon of 2 June 2002, the
observational record cannot document the conditions in the immediate vicinity of the fire,
nor provide sufficient spatial and temporal detail to associate the observed fire behavior
with specific atmospheric phenomena. The MM5 model described in section 2 is
employed to provide a more complete understanding of the sequence of events that
occurred during the fire. All of the model results discussed below are derived from a
simulation executed on a 4-km grid spacing model domain covering the northeastern
9
United States. The MM5 is initialized at 1200 UTC 1 June 2002 using NARR data for
initial conditions and boundary conditions.
Figure 9 shows the surface winds and relative humidity from the MM5
simulation, valid at 1800 UTC on 2 June. The simulated surface winds compare
favorably with observations (Fig. 3b). Northwesterly flow prevails across the MidAtlantic states with simulated wind speeds generally around 5 ms-1 across the region,
with stronger winds evident in eastern and southern NJ. The simulation depicts the the
weak dry front seen in the observations just off-shore by 1800 UTC in the simulation,
with wind shifts evident off the coast of northern NJ, NY, and Connecticut (CT). The
simulated surface relative humidity distribution agrees with the observed moisture
conditions in Fig. 3b, with dry air evident in southern NJ and MD. While surface air is
generally expected to be dry behind a cold front, it is noteworthy that isolated areas of
extremely low relative humidity (less than 30% in the simulation) have developed by
1800 UTC. While the synoptic-scale dry air can be attributed to synoptic forcing, the
local minimums in relative humidity near the fire location suggest that a local, mesoscale
forcing mechanism is responsible.
At 700 hPa, a broad area of extremely dry air is evident in the simulation (Fig.
10), with relative humidities below 5% in many locations across the Mid-Atlantic states.
This 700 hPa pool of dry air shows the horizontal extent of the dry feature evident above
750 hPa in the 0000 UTC 3 June observed sounding at OKX (Fig. 6b). While the origins
of this pool of dry air are explored in detail in Kaplan et al. (2008), for the purposes of
this discussion, the existence and location of the dry pool is more important than its
formation mechanisms. Since the model indicates that dry air was in place at 700 hPa
10
throughout this region, any process capable of bringing this dry air to the ground could
potentially generate weather conditions conducive to severe fire behavior. A Skew-T
diagram computed from the simulation at the fire location (39.9˚N, 74.1˚W) valid at 1800
UTC (Fig. 11) shows that that the simulated pool of dry air aloft at the fire location
extends from 850 hPa to the top of the atmosphere. Although the overall atmospheric
structure resembles the 0000 UTC sounding at OKX (Fig. 6b), the simulated sounding
suggests that a deeper layer of dry air is present at the fire location than at OKX, with the
driest air occurring between 800 and 650 hPa at the fire location.
A northwest-southeast vertical cross section (see Fig. 10 for the cross section
location) of simulated wind speed and relative humidity through the fire location
identifies two locations where deep (100-200 hPa) “shafts” of dry air overlay minimums
in surface relative humidity (Fig. 12). At 1700 UTC (Fig. 12a) one such shaft is located
directly above the fire, placing the surface relative humidity minimum directly over the
fire location. About 200 km to the northwest, coincident with the tallest Appalachian
peaks in this cross section, another shaft of locally low relative humidity is in place above
a somewhat weaker surface anomaly. The upstream feature appears to be generated by a
combination of orography-induced subsidence at the surface and downward vertical
advection of dry air due to the generation of mountain waves. The dry air at the fire
location, however, is not clearly the result of either of these processes, although waves
propagating downstream from the mountains and to the fire location can be seen in the
wind field at this time. By 1800 UTC (Fig. 12b), the dry shaft near the fire location has
propagated towards the southeast. Interestingly, a secondary relative humidity minimum
has formed at the ground, 20 km southeast of the fire, directly below the location of the
11
1700 UTC relative humidity minimum at 900 hPa (Fig. 12a). These analyses strongly
suggest that the surface air coupling with dry air aloft is associated with formation of
these surface-based relative humidity minimums.
Time series of simulated surface relative humidity (Fig. 14a), surface wind speed
(Fig. 14b), and PBL depth (Fig. 14c) at the fire location reveal that the onset of rapid PBL
growth between 1300 and 1400 UTC coincides with a rapid reduction in surface relative
humidity and an increase in surface wind speed. By 1700 UTC, when the initial attack
occurred on the fire (Table 1), the surface relative humidity reached its minimum value,
while wind speeds continued to climb through the afternoon. Time-height cross sections
of simulated relative humidity (Fig. 13a) and wind speed (Fig. 13b) at the fire location
show the vertical distribution of the simulated wind speed and relative humidity
throughout the day. These analyses confirm that the atmosphere 100-200 hPa above the
fire location was both considerably drier and exhibited higher wind speeds (momentum)
than at the ground. During the afternoon, two different episodes of vertical coupling
occur, one at 1700 UTC and the other at 2000 UTC. In the first episode, at 1700 UTC,
the dominant effect at the surface is drying, although a modest increase in surface wind
speed occurs as well. At 2000 UTC, the increase in wind speed is more substantial while
the drying signal is weaker. These episodic variations in wind speed in the simulation
occur during the same time period (1800 – 2000 UTC) when the firefighters reported
multiple wind shifts while fighting the fire.
Discussion and conclusion
12
Based on the observational and MM5 model analyses presented above, it appears
that coupling between the surface air and dry, high-momentum air aloft becomes possible
when the PBL deepens in the late morning and through the afternoon. There is also
evidence that high-frequency mountain waves generated by flow over the upstream
mountain range contribute to this coupling, once the PBL is sufficiently deep that
surface-based mixing processes can reach the pool of air aloft. The result is a series of
episodes wherein dry, high momentum air aloft mixes down to the ground and
contributes to sudden drying and highly variable winds (speed and direction) at the
surface. The transitory nature of the processes in this simulation suggests that coupling
between surface air and dry air aloft can have considerable spatial and temporal
variability for any given case.
Figure 15a shows simulated maps of the Haines index (HI) and the Ventilation
index (VI) for 1800 UTC 2 June. While the HI does “light up” over the mid-Atlantic
region at the time of interest, with values of 5 indicating a high potential for a blow-up
fire to occur in the absence of strong surface winds, the HI did not change dramatically
from the morning through the early afternoon (not shown). Thus, the HI failed to capture
the above-ground processes that accompanied the observed changes in surface conditions
that are hypothesized to have contributed to the observed extreme fire behavior during
this case. Figure 15b shows the VI at the same time at the HI. The VI, which is defined
as the mixed-layer depth multiplied by the mixed-layer average wind speed (e.g. Hardy et
al. 2002), also showed locally high values over southern NJ, as well as other locations in
the Mid-Atlantic states and New England. The VI is designed to assess smoke dispersion
from a wildland fire, and is not commonly employed as a fire behavior index. However,
13
the VI is sensitive to mixed layer depth and near-surface wind speeds, so the atmospheric
conditions aloft that are hypothesized to contribute to the observed fire behavior in this
case could lead to fine-scale variations in the index. But the simulated values of the VI
also fail to highlight the spatial and temporal variability in meteorological conditions that
accompanied the fire. Since the HI and the VI are the only commonly used operational
fire-weather indices that diagnose atmospheric conditions above the ground, it is clear
that new indices and diagnostics that are more sensitive to the meteorological processes
documented in this case could have provided useful guidance for preparing to fight this
fire.
The DT wildfire presented many challenges to firefighters and destroyed some
$400,000 worth of property. The tools available to the firefighters and the NWS were
helpful in effectively fighting the fire and preventing any loss of life or catastrophic loss
of property. Nevertheless, it is apparent that in this case some of the meteorological
processed that contributed to the observed fire behavior were not anticipated by the
available tools. The observations and MM5 simulations presented here suggest that
interactions between the PBL and deep pools of dry, high-momentum air aloft can
generate localized areas of dry, high momentum air at the surface. While the pool of
mid-tropospheric dry air is detectable in observations in this case, it is more
straightforward to detect and quantify in mesoscale simulations. These results are
supported by the findings of Mills (2003), Mills (2005a,b), and Zimet et al. (2007).
New mesoscale indices and diagnostics that diagnose specific types of weather
conditions aloft simulated by atmospheric models could provide additional predictive
power and early warnings of atmospheric conditions conducive to rapid fire growth.
14
There is an opportunity to develop indices and diagnostics that raise the “threat level”
when a pool of dry air is in place, and then can indicate localized areas of extreme danger
where the PBL is predicted to grow such that the dry air aloft will be mixed down to the
surface.
Case studies built around mesoscale model simulations, like the one presented
here, have the potential to radically alter our understanding of fire-weather interactions.
Formulating new indices and diagnostics based on a physical understanding of how a fire
modifies the atmosphere and how the atmosphere can impact fire behavior is an active
area of fire weather research. By conducting similar case studies for significant wildfires
both in the eastern region and throughout the United States, it is possible to understand
these interactions better and improve all facets of fire-atmosphere interaction prediction,
assessment, and understanding.
References
Brotak, E.A., and W. E. Reifsnyder, 1977: Predicting major wildland fire occurrence.
Fire Management Notes 38, 5-8.
Byram, G. M., 1954: Atmospheric conditions related to blowup fires. United States
Department of Agriculture Forest Service Station Paper 35, 31 pp.
Carlson, T. N., S. G. Benjamin, G. S. Forbes, and Y.-F. Li, 1983: Elevated mixed layers
in the severe storm environment - Conceptual model and case studies. Mon. Wea.
Rev., 111, 1453-1473.
Charney and Keyser, 2004: NWP conference paper on mixed-layer impact on fire
indices and diagnostics.
Chen, F., and J. Dudhia. 2001: Coupling an advanced land surface-hydrology model with
the Penn State/NCAR MM5 modeling system. Part I: model implementation and
sensitivity. Mon. Wea. Rev., 129, 569-585.
Davis, R. T., 1969: Atmospheric stability forecast and fire control. Fire Control Notes,
30, 3-14,15.
Fahnstock, G. R., 1965: Texas Forest Fires. United States Department of Agriculture
Forest Service Research General Technical Report, SO-16, 19 pp.
15
Grell, G. A., J. Dudhia, and D. R. Stauffer, 1995. A description of the fifth generation
Penn State/NCAR Mesoscale Model (MM5). NCAR Technical Note NCAR/TN398+STR. 122 p.
Heilman, W. E., B. E. Potter, J. J. Charney, and X. Bian, 2003: National Fire Plan—
Eastern Area Modeling Consortium: Research, product development, and
partnerships. 5thSymposium on Fire and Forest Meteorology, 17-20 November
2003, Orlando, FL, American Meteorological Society
Hardy, C.; Ottmar, R. D.; Peterson, J.; Core, J. 2001. Smoke management guide for
prescribed and wildland fire -- 2000 edition. PMS 420-2. NFES 1279. Boise, ID:
National Wildfire Coordination Group. 226 pp.
Janjic, Z. A., 1990: The step-mountain coordinate:Physical package. Mon. Wea. Rev.,
118, 1429-1443.
Kain, J. S., and J. M. Fritsch, 1990: A one-dimensional entrain-ing/detraining plume
model. J. Atmos. Sci., 47, 2784–2802.
Kaplan, M. L., C. Huang, Y.-L. Lin, and J. J. Charney, 2008: The development of
extremely dry surface air due to vertical exchanges under the exit region of a jet
streak. Submitted to Mon. Wea. Rev. (NOTE: 2nd revision is currently under
review, if the paper is not accepted by the time we submit this paper to
IJWF, I believe we can eliminate the references to this paper.)
Karyampudi, V. M., M. L. Kaplan, S. E. Koch, and R. Zamora, 1995: The influence of
the Rocky Mountains on the 13-14 April 1986 severe weather outbreak. Part I:
Mesoscale lee cyclogenesis and its relationship to severe weather and dust storms.
Mon. Wea. Rev., 123, 1394-1422.
Keetch, J. J., and G. M. Byram, 1968: A drought index for forest fire control. USDA
Forest Service Research Paper SE-38, Southeastern Forest Experiment Station,
Asheville, NC, 33 pp
Lansing, L., 1939: Weather preceding forest fires in New Hampshire. Bul. Amer. Met.
Soc., 20, 10-26.
Mesinger, F., G. Dimego, E. Kalnay, K. Mitchell, P. C. Sharfran, W. Ebisuzaki, D. Jovic,
E. Rodgers,E. Berbery, M. B. Ek, Y. Fan, R. Grumbine, W.Higgins, H. Li, Y. Lin,
G. Manikin, D. Parrish,W. Shi, 2006: North American Regional Reanalysis. Bull.
Amer. Meteor. Soc. 87. 343-360.
Mills, G.A. and Pendlebury, S., 2003. Processes leading to a incident at Hobart Airport.
Aust. Meteor.Mag. 52, 171-188.
Mills, G.A., 2005. A re-examination of the synoptic and mesoscale Ash Wednesday
1983. Aust. Meteor. Mag. 54, 35-55.
Mills, G.A., 2005. On the sub-synoptic scale meteorology of two extreme fire weather
days during the Eastern Australian fires of January 2003. Aust. Meteor. Mag.
Mlawer, E. J., S. J. Taubman, P. D. Brown, M. J. Iacono, and S. A. Clough, 1997:
Radiative transfer for inhomogeneous atmospheres: RRTM, a validated
correlated-k model for the longwave. J. Geophys. Res., 102, 4353–4356.
New Jersey State Forest Fire Service (NJFFS), 2003: Division forest firewardens report
of large fire and problem fire analysis: Jakes Branch Wildfire B06-02-02 (2002),
20 pp.
16
Simard, A. J., J. E. Eenigenburg, and S. L. Hobrla, 1987: Predicting extreme fire
potential. Proceedings of the Ninth Conference on Fire and Forest Meteorology,
American Meteorological Society, San Diego, California, April 21-24, 1987, 148157.
Zhong, S., H.-J. In, X. Bian, J. Charney, W. Heilman, and B. Potter, 2005: Evaluation of
real-time high-resolution MM5 predictions over the Great Lakes region. Wea.
Forecasting, 20, 63–81.
Zimet, T., J.E. Martin and B.E. Potter, 2007: The influence of an upper-level frontal
zone on the Mack Lake wildfire environment. Meteorological Applications, in
press.
17
Download