Вы находитесь на странице: 1из 23

A Diagnostic Case Study Analysis of A Mesoscale Snowband Over

North Texas

Jonathan C. Whitehead
School of Meteorology
University of Oklahoma

ABSTRACT
On March 6, 2008 a narrow band of heavy snow fell across portions of North Texas. This
paper is a case study analysis into the processes that caused this mesoscale snowband. At the
surface, low pressure was centered over eastern Texas with a shortwave trough approaching
from the west at midlevels. Throughout the day midtropospheric frontogenesis increased as
warm, moist air became trapped in the trough of warm air aloft (trowal) to the west and
northwest of the surface cyclone. A vertical cross-section, taken normal to the 850-300 hPa
thicknesses, revealed folded θe surfaces indicative of convective instability within a region of
small positive equivalent potential vorticity. Finally, sounding analysis shows a nearly moist
adiabatic profile, an isothermal temperature profile in the PBL, and a distinct col between 600
and 700 hPa; all of which points to an atmosphere conducive for heavy banded snowfall.

1. Introduction

On March 6, 2008 a heavy band of snow fell across parts of North Texas

stretching from Stephens Co. northeast into Grayson Co. National Weather Service

(NWS) estimated snowfall totals were as high as nine inches or more with this swath of

heavy snow (Fig. 1). The half-width of the snowband (the distance from max snowfall

totals to half that value) was approximately 65 km. Radar reflectivity values were as high

as 50 dBZ within the band of heavy snowfall (Fig. 2). With this background info in mind,

I propose a diagnostic case study analysis into the atmospheric processes that helped to

force and focus this narrow region of heavy snowfall. The primary scientific goal

representing the core intellectual merit of the proposed research is toward understanding

the organization of the extensive mesoscale snowband. The motivation behind this case

study stems from the continued challenge mesoscale snowbands present to operational

meteorologist. Banacos (2003) makes the point that the spatial location and duration of

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
these heavy snowbands is often difficult to predict accurately. It is important to note that

a number of case studies of heavy snowfall from mesoscale bands in the plains have been

published over the past 20 years (i.e., Moore et al. 2005; Marwitz and Toth 1993; Trapp

et al. 2001; Bennetts and Hoskins 1979). Despite the past studies there is still a lot left to

learn. The main challenge to this heavy snow event is to explain the length and breadth of

the heavy snowfall as well as thundersnow in the presents of inherently weak surface

cyclones (Moore et al. 2005). Toward that end, the proposed research will attempt to

identify the ingredients that came together resulting in this narrow corridor of heavy

snowfall.

The broader impacts of the proposed research lie primarily in the possibilities of

its operational applications. While it is virtually impossible to make an accurate forecast

more then two days out based on diagnostic data alone, it is hoped that a thorough

understanding of the processes involved in this event will be of use to forecasters so that

they may be able to identify these processes based strictly off of diagnostic data, and,

therefore, make more accurate one to two day forecast based off the research.

2. Background

a. Conveyor belts

To adequately understand the processes involved in creating a mesoscale

snowband, it is essential that one have a thorough understanding of the Norwegian

cyclone model and how various airstreams within the cyclone interact in enhancing

snowband formation. Research by Harold (1973), Carlson (1980), and Danielson (1964)

has identified three major airstreams associated with cyclogenesis, termed the warm,

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
cold, and dry conveyor belts. These belts have been shown to directly influence and

dictate the organization of precipitation attending extratropical cyclones (ETC) (Fig. 3). It

is the 3D interaction of these three airstreams in the vicinity of the ETC that can lead to a

favorable environment for the formation of heavy banded precipitation (Nicosia and

Grumm, 1999).

The interaction of these conveyor belts provides the moisture, instability, and lift.

Moisture and instability are provided by the cyclonically curving branch of the warm

conveyor belt (WCB) northwest of the surface low, lift is associated with midlevel

frontogenesis, and the dry conveyor belt (DCB) is associated with enhancing the

instability. Martin (1998 a,b) has shown that the bifurcation of both the WCB and cold

conveyor belt (CCB) north of the warm front creates a deformation zone that acts on the

potential temperature gradient through stretching and shearing, resulting in midlevel

frontogenesis to the northwest of the surface low. Additionally, the DCB acts to

destabilize the atmosphere through the advection of dry air over low-level moist air

(Danielson, 1964).

b. Instability

While we’re all familiar with the classic severe weather instability parameter of

CAPE, there are also some measures of instability that are strictly used for winter

weather convection. Bennetts & Hoskins (1979) were among the first to show how

frontal rainbands might be explained by the presence of conditional symmetric instability

(CSI), a condition wherein the atmosphere is convectively stable (i.e., equivalent

potential temperature, θe, increasing with height) and intertially stable (geostrophic

absolute vorticity greater than zero, ηg>0), yet is unstable to slantwise ascent (Moore et
Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120
David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
al., 2005). Emanuel (1985) has shown that the upward vertical motion branch of a

frontal-scale circulation in the presence of CSI or even weak symmetric stability (WSS)

is both enhanced and contracted. Moore and Lambert (1993) developed a 2D form of

assessing regions of CSI within a cross-sectional plane known as equivalent potential

vorticity (EPV). McCann (1995) took this idea and applied it to a 3D form showing that

CSI can be diagnosed in a region of negative EPV (EPV<0) that tends to form in a

saturated environment within which the vertical wind shear is strong and convective

stability is weak. However, when EPV≤0.25, the environment is conducive to weak

symmetric stability and single band formation is preferred (Schumacher, 2003).

3. Methodology

a. Instability

In diagnosing regions of CSI, I used the Moore and Lambert (1993) cross-

sectional approach previously mentioned. In this approach, CSI is evaluated in the cross-

section taken normal to the mid-tropospheric thermal wind by displaying lines of constant

Mg and θe. Emanuel (1983) defines Mg as the absolute geostrophic momentum. In order

to produce the required fields of Mg and θe, I used objectively analyzed fields of θe, ug,

and vg (geostrophic wind components) at 11 levels from 1000 to 100 hPa. Upon choosing

a northern and southern point for the cross-section, taken normal to the 850-300 hPa

thickness, values of θe and the geostrophic wind component normal to the cross-section

were interpolated to the line of line of the cross-section.

Moore and Lambert (1993) evaluated CSI by qualitatively comparing the slope of

the θe surfaces with that of the Mg surfaces. CSI is diagnosed in these regions where the
Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120
David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Mg surfaces are “flatter” (more horizontal) than the θe surfaces. It can be shown that in

those regions, a parcel is stable with respect to slantwise ascent. It is important to note

that when diagnosing CSI, relative humidities (RH) should exceed 80% (Bennetts and

Sharp, 1982).

In computing EPV, Moore and Lambert (1993) expanded the EPV equation

following Martin et al. (1992) to yield a 2D form:

(1)
(A) (B)
Term (A) represents the contribution to EPV from the vertical wind shear and the

horizontal temperature gradient. Therefore, when term (A) is large, EPV becomes

negative, indicating CSI or CI (convective instability). Term (B) represents absolute

vorticity and a measure of convective stability. Term (B) is generally positive by

definition of Mg. Since the positive x direction in these sections points toward warmer

air, EPV surfaces slope down and term (B) will be greater than zero. Therefore, the net

result of term (A) multiplied by term (B) will be negative and made more so by a stronger

horizontal θe gradient or vertical wind shear (Moore and Lambert, 1993).

b. Frontogenesis

When investigating regions of CSI, it is ironic how often you will find the

presence of frontogenesis within regions of CSI. There is a reason for this. According to

Nicosia and Grumm (1999) and Moore et al. (2005), in a region of frontogenesis, the

gradient of potential temperature increases with time. By the thermal wind relationship,

this requires an increase in the geostrophic wind shear, which results in differential

moisture advection that steepens the vertical slope of θe isentropes. The subsequent

weakening of the convective instability, together with a strengthening of the vertical wind
Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120
David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
shear, results in a reduction of EPV, often leading to CSI. What’s important here is that

these processes can occur in or near a region of weak cyclogenesis quite a distance away

from the surface low center.

Frontogenesis is the lagrangian time rate of change of the magnitude of the

horizontal potential temperature gradient. Petterssen (1956) expresses this as:

1
F= | ∇θ | [ Defr cos(2β)-Div] (2)
2

where ∇θ is the potential temperature gradient, Defr is the resultant deformation, β is the

angle between the isentropes and the axis of dilation, and Div is divergence.

4. Data

The case under examination covers the period 1200 UTC 6 March—0000 UTC 7

March 2008. Hourly surface data came from the Plymouth St. archive webpage. Standard

upper-air analysis came from the Storm Prediction Center (SPC) map archive webpage.

Fields of EPVg and frontogenesis were collected via the SPC mesoanalysis archive

webpage. Derived fields of observed surface and upper-air parameters were computed

from the RUC II using the General Meteorological Package (GEMPAK) available to

academic institutions. Level III radar data were obtained using the Weatherscope®

software available through the Oklahoma Climatological Survey (OCS) website.

5. Diagnostic Analysis

a. Surface

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Surface analysis for the period 1500 UTC—2100 UTC 6 March shows a quasi-

stationary surface cyclone over eastern Texas (TX) (Fig 4 a-c). This is further evidenced

by 19Z MSLP and surface wind analysis from the SPC (Fig. 5). Note the north-south

trough axis in central TX and the southwest-northeast oriented ridge axis from the TX

Gulf Coast into southwestern Arkansas. A closer inspection of surface observations

shows temperatures at 15Z behind the front in the 40s along the I-35 corridor, dropping to

the 30s in central and west TX. Snow is already being reported by stations across

northwest TX into the Panhandle. By 2100 UTC the temperature gradient along the cold

front has increased significantly with 60s right along the coast, dropping to the 40s about

50 miles inland, and down to the 30s north and west of Waco. At this time snow has

begun to fall across much of N. TX. It is important to note that this snow fell north and

west of the surface cyclone, which fits well with the conveyor belt model discussed in

section 2.

b. Upper-level flow

Initially at 1200 UTC 6 March at 850 hPa, a shortwave (hereafter, s/w) trough

was located over west TX with a strong thermal gradient along and to the east of the

trough axis. Over the next 12 hours (Fig. 6 a,d), the s/w trough moves very little. Initially

at 500 hPa, a longwave trough covers much of the continental U.S. (hereafter, CONUS)

with a positively-tilted trough axis from North Dakota to New Mexico. Over subsequent

time periods (Fig. 6 b,e), the trough axis shifts east into west TX. A 300 hPa isotach

analysis for the two time periods (Fig. 6 c,f) reveals a distinctly amplified flow over the

CONUS with a split flow regime over the eastern half of the nation. The main polar jet is

analyzed over the Midwest with an embedded 125kt jet streak over northeastern Indiana,
Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120
David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
northwest Ohio, and southeast Michigan. At the same time a 100kt jet streak is

propogating around the base of the trough along the Baja Peninsula.

The critical aspect in the upper-air analysis is the dramatic evolution of the Polar

jet over the Midwest and the secondary jet across the Baja. During the 12hr period, the

Polar jet over the Midwest has expanded considerably and back-built to the Missouri

River Valley. The jet over the Baja has shifted eastward into northern Mexico. The

orientation of these two jet streaks has quite possibly setup a coupled jet scenario over N

TX. The back-building jet is due in part to the positive feedback processes of the

mesoscale snowband. This process involves the large amount of latent heat release given

off by the snowband due to the presence of high instability (shown later). The latent heat

release creates a meso-high aloft and the ageostrophic response from the meso-high

enhances the southwesterly flow over the region, and, therefore, the Midwest jet appears

to back-build southwestward.

c. Midlevel frontogenesis and EPV

Plane views of layer-averaged frontogenesis for the 850-700 hPa (Fig. 7) show a

frontogenetic maximum that consistently moves slowly to the east-northeast from west-

central TX to north-central TX. This frontogenetic maximum is located well to the north

of the cold front and west of the surface cyclone. Plane view plots of layer-averaged

saturated EPVg for 850-700 hPa for the 1500-2100 UTC time period (Fig. 8) consistently

reveal a region of negative to slightly positive EPVg over N TX. This would indicate

either CSI, CI, or WSS is the dominate instability type.

Recall the Moore and Lambert approach to diagnosing CSI in regions of negative

EPVg in which cross-sections were taken normal to the 850-300 hPa thickness values. For
Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120
David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
the 6 March event, a vertical cross-section Mg and θe was taken normal to the 18Z 850-

300 hPa thicknesses (Fig. 9) from TYR to SPS (black line). This particular line and time

was chosen since the KFWS radar at 18Z (Fig. 2) reported a well-defined snowband

normal to the cross-sectional line. The outline area on the cross-section (Fig. 10) is a

region where the θe surfaces are folded over, indicative of CI. Moore and Lambert (1993)

note that “this folded region is a region of CI created from warm, moist air riding over a

frontal boundary. Areas with cold boundary layer temperatures (low θe values) with CI

aloft are susceptible to elevated convection given upward vertical motion and near

saturated conditions”. This condition of saturation is proven by inspection of the 18Z

FWD sounding (Fig. 11).

Analysis of the 18Z FWD sounding reveals several features that would point to

heavy snow. First, there is a deep, moist nearly isothermal lapse rate from approximately

950 to 800 hPa. According to Moore et al. (2005) “such isothermal lapse rates are due to

atmospheric cooling as ice crystals aloft fall into a warm layer with temperatures over

0°C. This cooling can result in a relatively deep (often up to 1 Km) isothermal layer at or

below 0°C. This saturated isothermal layer is often associated with a mesoscale indirect

thermal circulation which enhances precipitation amounts near rain-snow boundaries.” A

second feature of the sounding is that above the isothermal layer there is a deep, moist-

adiabatic lapse rate extending to about 200 hPa. This shows that there was a deep layer of

moisture with cloud temperatures well below -5°C, so that ice crystals were plentiful to

see the warmer, supercooled layer below. A third and final feature of the sounding is that

the vertical wind profile reveals a distinct col region between 600 and 700 hPa. Banacos

(2003) has noted that “the best scenario from a banding perspective would appear to be

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
strong deformation (and convergence) with minimum translation or vorticity, such that a

col point exists in a system-relative and an absolute sense, creating a very favorable

environment not only for banding, but also for long-lived bands affecting one specific

area.”

6. Conclusion

A comprehensive case study analysis has been presented to document the


ingredients that came together to focus a narrow and heavy mesoscale snowband across
N TX. This case demonstrated characteristics represented in the conceptual models from
Moore et al. (2005) and Moore and Lambert (1993). In the Moore et al. (2005) model, the
synergistic interaction of the major conveyor belts came together to create a mesoscale
region of enhanced lift, instability, and deep moisture supportive of heavy snow. In the
Moore and Lambert model, a cross section through a region of EPVg close to zero
revealed a CI regime that, together with a saturated vertical profile, was supportive of an
upright elevated convective snowband. It is hoped that analysis of these key features can
be used in an operational sense to accurately forecast the length and breadth of heavy
snowfall in banded environments.

Acknowledgments. The author would like to thank Nick Hampshire of the


NWSFO-FWD for his guidance in getting the author started on the research. The author
would also like to thank Greg Carbin of the SPC for providing the GEMPAK scripts and
Patrick Marsh for his assistance in generating the images from the scripts. A final thanks
goes to Dr. William Beasley who provided the author with continued positive support and
guidance throughout the capstone process.

REFERENCES

Banacos, P.C., 2003: Short-range prediction of banded precipitation associated with


deformation and frontogenetic forcing. Preprints, 10th Conf. on Mesoscale
Processes, Portland, OR, Amer. Meteor. Soc., CD-ROM, P1.7.

Bennetts, D. A., and B. J. Hoskins, 1979: Conditional symmetric instability—A possible


explanation for frontal rainbands. Quart. J. Roy. Meteor. Soc., 105, 945-962.

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
______, and J. D. Sharp, 1982: The relevance of conditional symmetric instability to the
prediction of mesoscale frontal rainbands. Quart. J. Roy. Meteor. Soc., 108, 595-
602.

Carlson, T. N., 1980: Airflow through midlatitude cyclones and the comma cloud pattern.
Mon. Wea. Rev., 108, 1498-1509.

Danielson, E. F., 1964: Project Springfield Report. DASA Rep. 1517, Defense Atomic
Support Agency, 97 pp. [NTIS AD-607980.]

Emanuel, K. A., 1983: The Lagrangian parcel dynamics of moist symmetric stability. J.
Atmos. Sci., 40, 2368-2376.

______, 1985: Frontal circulations in the presence of small moist symmetric instability. J.
Atmos. Sci., 42, 1062-1071.

Harold, T. W., 1973: Mechanism influencing the distribution of precipitation within


baroclinic disturbances. Quart. J. Roy Meteor. Soc., 99, 232-251.

Martin, J. E., 1998a: The structure and evolution of a continental winter cyclone. Part I:
Frontal structure and the occlusion process. Mon. Wea. Rev., 126, 303-328.

______, 1998b: The structure and evolution of a continental winter cyclone. Part II:
Frontal forcing of an extreme snow event. Mon. Wea. Rev., 126, 329-348.

Marwitz, J., and J. Toth, 1993: A case study of heavy snowfall in Oklahoma. Mon. Wea.
Rev., 121, 648-660.

McCann, D. W., 1995: Three-dimensional computations of equivalent potential vorticity.


Wea. Forecasting, 10, 798-802.

Moore, J. T., and T. E. Lambert, 1993: The use of equivalent potential vorticity to
diagnose regions of conditional symmetric instability. Wea. Forecasting, 8, 301-
308.

______, C. E. Graves, S. NG, and J. L. Smith, 2005: A process-oriented methodology


toward understanding the organization of an extensive mesoscale snowband: A
diagnostic case study of 4-5 December 1999. Wea. Forecasting, 20, 35-50.

Nicosia, D. J., and R. H. Grumm, 1999: Mesoscale band formation in three major
northeastern United States snowstorms. Wea. Forecasting, 14, 346-368.

Petterssen, S., 1956: Weather Analysis and Forecasting. Vol. 1. McGraw-Hill, 428 pp.

Schumacher, P. N., 2003: An example of forecasting mesoscale bands in an operational


environment. Preprints, 10th Conf. on Mesoscale Processes, Portland, OR, Amer.
Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120
David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Meteor. Soc., P1.11. [Available online at
http://ams.confex.com/ams/pdfpapers/62199.pdf ]

Trapp, R. J., D. M. Schultz, A. V. Ryzhkov, and R. L. Holle, 2001: Multiscale structure


and evolution of an Oklahoma winter precipitation event. Mon. Wea. Rev., 129,
486-501.

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figures

Figure 1. NWS estimated snowfall totals from 6 March 2008. (NWS)

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 2. KFWS WSR-88D Level III composite reflectivity for 1800 UTC 6 March 2008
(OCS)

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 3. Conveyor belts associated with extratropical cyclones. (A color version of this
schematic diagram is available online: www.eas.slu.edu/CIPS/Publications.html )

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 4. Surface observations for (a) 1500 UTC, (b) 1800 UTC, and (c) 2100 UTC 6
March 2008. (PSU)

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 5. Mesoanalysis of MSLP and surface wind for 1900 UTC 6 March 2008. (SPC)

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
12 UTC 6 March 2008 00 UTC 7 March 2008

Figure 6. Subjective analysis of upper-level pressure surfaces: (a) 850 hPa for 1200 UTC 6 March 2008;
solid lines are isohypses in dm and dashed lines are isotherms in °C; (b) 500 hPa for 1200 UTC 6 March
2008; solid lines are isohypses in dm and dashed lines are isotherms in °C; (c) 300 hPa for 1200 UTC 6
March 2008; solid lines are streamlines, fill regions are isotachs in kt, and yellow lines are divergence; (d)
same as in (a) except for 0000 UTC 7 March 2008; (e) same as in (b) except for 0000 UTC 7 March 2008;
and (f) same as in (c) except for 0000 UTC 7 March 2008. (SPC)

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 7. Average frontogenesis for the 850-700 hPa layer for (a) 1500 UTC, (b) 1800
UTC, and (c) 2100 UTC 6 March 2008. Solid lines are isohypses in dm, dashed lines are
isotherms in °C, and fill region is frontogenesis. (SPC)

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 8. Average saturated EPVg in units of PVU for the layer 800-750 hPa for (a) 1500
UTC, (b) 1800 UTC, and (c) 2100 UTC 6 March 2008. Solid red lines are 850 hPa
frontogenesis and shaded region is EPVg. (SPC)
Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120
David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 9. The 850-300 hPa thickness (gpm) for 1800 UTC 6 March 2008. Black line
depicts the position of cross-section shown in Figure 10.

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 10. Vertical cross-section of absolute geostrophic momentum (Mg, dashed lines)
and equivalent potential temperature (θe, solid red lines) along the cross-section in Fig. 9
for 1800 UTC 6 March 2008. Circled region denotes region of CI.

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net
Figure 11. FWD sounding for 1800 UTC 6 March 2008. (NWS)

Corresponding author address: Jonathan C. Whitehead, OU School of Meteorology, 120


David L. Boren Blvd., Norman, OK 73072
Email: ouweathersooner@earthlink.net

Вам также может понравиться