20.01.2015 Views

Environmental Control of Cloud-to-Ground Lightning Polarity in ...

Environmental Control of Cloud-to-Ground Lightning Polarity in ...

Environmental Control of Cloud-to-Ground Lightning Polarity in ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

APRIL 2007 C A R E Y A N D B U F F A L O 1327<br />

<strong>Environmental</strong> <strong>Control</strong> <strong>of</strong> <strong>Cloud</strong>-<strong>to</strong>-<strong>Ground</strong> <strong>Lightn<strong>in</strong>g</strong> <strong>Polarity</strong> <strong>in</strong> Severe S<strong>to</strong>rms<br />

LAWRENCE D. CAREY AND KURT M. BUFFALO*<br />

Department <strong>of</strong> Atmospheric Sciences, Texas A&M University, College Station, Texas<br />

(Manuscript received 22 February 2006, <strong>in</strong> f<strong>in</strong>al form 27 July 2006)<br />

ABSTRACT<br />

In this study, it is hypothesized that the mesoscale environment can <strong>in</strong>directly control the cloud-<strong>to</strong>-ground<br />

(CG) lightn<strong>in</strong>g polarity <strong>of</strong> severe s<strong>to</strong>rms by directly affect<strong>in</strong>g their structural, dynamical, and microphysical<br />

properties, which <strong>in</strong> turn directly control cloud electrification and ground flash polarity. A more specific<br />

hypothesis, which has been supported by past observational and labora<strong>to</strong>ry charg<strong>in</strong>g studies, suggests that<br />

broad, strong updrafts and associated large liquid water contents <strong>in</strong> severe s<strong>to</strong>rms lead <strong>to</strong> the generation <strong>of</strong><br />

an <strong>in</strong>verted charge structure and enhanced CG lightn<strong>in</strong>g production. The corollary is that environmental<br />

conditions favor<strong>in</strong>g these k<strong>in</strong>ematic and microphysical characteristics should support severe s<strong>to</strong>rms generat<strong>in</strong>g<br />

an anomalously high (25%) percentage <strong>of</strong> CG lightn<strong>in</strong>g (i.e., positive s<strong>to</strong>rms) while environmental<br />

conditions relatively less favorable should susta<strong>in</strong> s<strong>to</strong>rms characterized by a typical (25%) percentage <strong>of</strong><br />

CG lightn<strong>in</strong>g (i.e., negative s<strong>to</strong>rms). Forty-eight <strong>in</strong>flow proximity sound<strong>in</strong>gs were analyzed <strong>to</strong> characterize<br />

the environment <strong>of</strong> n<strong>in</strong>e dist<strong>in</strong>ct mesoscale regions <strong>of</strong> severe s<strong>to</strong>rms (4 positive and 5 negative) on 6 days<br />

dur<strong>in</strong>g May–June 2002 over the central United States. This analysis clearly demonstrated significant and<br />

systematic differences <strong>in</strong> the mesoscale environments <strong>of</strong> positive and negative s<strong>to</strong>rms, which were consistent<br />

with the stated hypothesis. When compared <strong>to</strong> negative s<strong>to</strong>rms, positive s<strong>to</strong>rms occurred <strong>in</strong> environments<br />

associated with a drier low <strong>to</strong> midtroposphere, higher cloud-base height, smaller warm cloud depth, stronger<br />

conditional <strong>in</strong>stability, larger 0–3 km AGL w<strong>in</strong>d shear, stronger 0–2 km AGL s<strong>to</strong>rm relative w<strong>in</strong>d speed, and<br />

larger buoyancy <strong>in</strong> the mixed-phase zone, at a statistically significant level. Differences <strong>in</strong> the warm cloud<br />

depth <strong>of</strong> positive and negative s<strong>to</strong>rms were by far the most dramatic, suggest<strong>in</strong>g an important role for this<br />

parameter <strong>in</strong> controll<strong>in</strong>g CG lightn<strong>in</strong>g polarity. In this study, strong correlations between the mesoscale<br />

environment and CG lightn<strong>in</strong>g polarity were demonstrated. However, causality could not be verified due <strong>to</strong><br />

a lack <strong>of</strong> <strong>in</strong> situ observations <strong>to</strong> confirm the hypothesized microphysical, dynamical, and electrical responses<br />

<strong>to</strong> variations <strong>in</strong> environmental conditions that ultimately determ<strong>in</strong>ed the dom<strong>in</strong>ant CG polarity. Future<br />

observational field programs and cloud model<strong>in</strong>g studies should focus on these critical <strong>in</strong>termediary processes.<br />

1. Introduction<br />

Although the overwhelm<strong>in</strong>g majority (i.e., about<br />

90%) <strong>of</strong> ground flashes lower net negative charge<br />

across the contiguous United States (CONUS; Orville<br />

and Huff<strong>in</strong>es 2001), a few severe s<strong>to</strong>rms can generate<br />

positive cloud-<strong>to</strong>-ground (CG) flash rates, densities,<br />

and percentages comparable <strong>to</strong> those typically observed<br />

for negative cloud-<strong>to</strong>-ground (CG) flashes <strong>in</strong><br />

* Current affiliation: NOAA/National Weather Service,<br />

Weather Forecast Office, Hast<strong>in</strong>gs, Nebraska.<br />

Correspond<strong>in</strong>g author address: Dr. Lawrence D. Carey, Department<br />

<strong>of</strong> Atmospheric Sciences, Texas A&M University, 3150<br />

TAMU, College Station, TX 77843-3150.<br />

E-mail: larry_carey@tamu.edu<br />

active thunders<strong>to</strong>rms (e.g., MacGorman and Burgess<br />

1994; S<strong>to</strong>lzenburg 1994; Carey and Rutledge 1998; Lang<br />

and Rutledge 2002; Carey et al. 2003a).<br />

The large majority (about 80%) <strong>of</strong> warm-season severe<br />

s<strong>to</strong>rms throughout the CONUS generate mostly<br />

(75%) negative ground flashes (i.e., so-called negative<br />

s<strong>to</strong>rms) while a m<strong>in</strong>ority (about 20%) produce an<br />

anomalously high (25%) percentage <strong>of</strong> positive<br />

ground flashes (i.e., so-called positive s<strong>to</strong>rms; Carey et<br />

al. 2003b, hereafter CRP03). 1 The frequency <strong>of</strong> these<br />

1 The def<strong>in</strong>ition <strong>of</strong> anomalous positive or positive s<strong>to</strong>rms is<br />

somewhat arbitrary <strong>in</strong> the literature with values <strong>of</strong> CG fraction<br />

rang<strong>in</strong>g from 25% <strong>to</strong> 50%. CRP03 found the geographic distribution<br />

<strong>of</strong> severe s<strong>to</strong>rms with this range <strong>of</strong> CG fraction <strong>to</strong> be<br />

identical. Hence, we choose <strong>to</strong> use the less restrictive def<strong>in</strong>ition <strong>of</strong><br />

25% CG fraction.<br />

DOI: 10.1175/MWR3361.1<br />

© 2007 American Meteorological Society


1328 MONTHLY WEATHER REVIEW VOLUME 135<br />

FIG. 1. Percentage (color shaded as shown) <strong>of</strong> severe s<strong>to</strong>rm reports associated with 25% CG lightn<strong>in</strong>g dur<strong>in</strong>g the warm season<br />

(April–September) dur<strong>in</strong>g 1989–98. The boxed region (dotted) <strong>in</strong>dicates the IHOP_2002 experimental doma<strong>in</strong>. The ellipse (dashed)<br />

marks the area <strong>of</strong> <strong>in</strong>terest for this study where both positive and negative polarity s<strong>to</strong>rms are common. The significant overlap between<br />

the two regions highlights the suitability <strong>of</strong> the IHOP_2002 dataset for this study. (Adapted from CRP03.)<br />

“anomalous” positive s<strong>to</strong>rms varies regionally (Fig. 1).<br />

Positive s<strong>to</strong>rms account for less than 10%–20% <strong>of</strong> all<br />

warm-season severe s<strong>to</strong>rms <strong>in</strong> the eastern and southern<br />

United States. In the central and northern pla<strong>in</strong>s<br />

from the Texas panhandle northeastward <strong>to</strong> M<strong>in</strong>nesota,<br />

30%–90% <strong>of</strong> all severe s<strong>to</strong>rms are positive<br />

s<strong>to</strong>rms.<br />

The geographic preference <strong>of</strong> positive s<strong>to</strong>rms for the<br />

central and northern pla<strong>in</strong>s over the eastern United<br />

States and southern pla<strong>in</strong>s suggests that such s<strong>to</strong>rms<br />

may be l<strong>in</strong>ked <strong>to</strong> specific meteorological conditions that<br />

are more prevalent <strong>in</strong> the favored regions. Several past<br />

studies have noted that severe s<strong>to</strong>rms pass<strong>in</strong>g through<br />

similar mesoscale regions on a given day exhibit similar<br />

CG lightn<strong>in</strong>g behavior (Branick and Doswell 1992;<br />

MacGorman and Burgess 1994; Smith et al. 2000), lead<strong>in</strong>g<br />

<strong>to</strong> the hypothesis that the local mesoscale environment<br />

<strong>in</strong>directly <strong>in</strong>fluences CG lightn<strong>in</strong>g polarity by directly<br />

controll<strong>in</strong>g s<strong>to</strong>rm structure, dynamics, and microphysics,<br />

which <strong>in</strong> turn control s<strong>to</strong>rm electrification (e.g.,<br />

MacGorman and Burgess 1994; Gilmore and Wicker<br />

2002, hereafter GW02; CRP03; Williams et al. 2005,<br />

hereafter W05).<br />

2. Background<br />

a. Thunders<strong>to</strong>rm charg<strong>in</strong>g and cloud-<strong>to</strong>-ground<br />

lightn<strong>in</strong>g polarity<br />

The ma<strong>in</strong> electrical dipole present <strong>in</strong> most thunders<strong>to</strong>rms<br />

(i.e., negative s<strong>to</strong>rms) is composed <strong>of</strong> an upperlevel<br />

positive over a low-level negative charge (Wilson<br />

1920). Of course, it is now well known that the electrical<br />

structure <strong>in</strong> deep convection is more complex with<br />

three (tripole upper positive, lower negative, and<br />

lower positive charges; Williams 1989) <strong>to</strong> four or five


APRIL 2007 C A R E Y A N D B U F F A L O 1329<br />

significant layers <strong>of</strong> charge <strong>in</strong> the vertical (S<strong>to</strong>lzenburg<br />

et al. 1998).<br />

Recent observations with balloon-borne electric field<br />

mills (Rust and MacGorman 2002) and (very high frequency)<br />

VHF-based <strong>to</strong>tal lightn<strong>in</strong>g mapp<strong>in</strong>g networks<br />

(Lang et al. 2004; Wiens et al. 2005) suggest that positive<br />

s<strong>to</strong>rms are characterized by an <strong>in</strong>verted-dipole or<br />

<strong>in</strong>verted-polarity vertical electric field structure, mean<strong>in</strong>g<br />

that the vertical configuration <strong>of</strong> the <strong>in</strong>ferred charge<br />

polarity structure is nearly opposite <strong>to</strong> the arrangement<br />

found <strong>in</strong> the more typical negative s<strong>to</strong>rms. In <strong>in</strong>vertedpolarity<br />

s<strong>to</strong>rms, the lowest significant charge layer <strong>of</strong><br />

the thunders<strong>to</strong>rm tripole is negative, the midlevel<br />

charge is positive, and the ma<strong>in</strong> upper-level charge is<br />

negative (i.e., opposite <strong>to</strong> the conventional tripole). 2 In<br />

one hypothesis (e.g., GW02; CRP03; W05), broad, <strong>in</strong>tense<br />

updrafts and associated high liquid water contents<br />

<strong>in</strong> positive s<strong>to</strong>rms lead <strong>to</strong> positive charg<strong>in</strong>g <strong>of</strong> graupel<br />

and hail and negative charg<strong>in</strong>g <strong>of</strong> ice crystals dur<strong>in</strong>g<br />

rebound<strong>in</strong>g collisions via the non<strong>in</strong>ductive charg<strong>in</strong>g<br />

(NIC) mechanism (e.g., Takahashi 1978; Jayaratne et<br />

al. 1983; Saunders et al. 1991; Saunders and Peck 1998).<br />

The differential fall speeds <strong>of</strong> ice particles and an accelerat<strong>in</strong>g<br />

updraft result <strong>in</strong> the s<strong>to</strong>rm-scale separation<br />

<strong>of</strong> the positively charged graupel and hail and the negatively<br />

charged ice crystals and the formation <strong>of</strong> an <strong>in</strong>verted<br />

dipole (i.e., negative over positive charge). This<br />

suggestion is sometimes broadly referred <strong>to</strong> as the “<strong>in</strong>verted-dipole”<br />

hypothesis. The more common NIC<br />

process, which is the negative charg<strong>in</strong>g <strong>of</strong> graupel and<br />

hail and positive charg<strong>in</strong>g <strong>of</strong> ice crystals, is hypothesized<br />

<strong>to</strong> occur <strong>in</strong> more typical s<strong>to</strong>rms with moderate<br />

updrafts and liquid water contents, result<strong>in</strong>g <strong>in</strong> the typical<br />

thunders<strong>to</strong>rm dipole and CG lightn<strong>in</strong>g. The tilted<br />

dipole hypothesis (Brook et al. 1982) suggests that the<br />

upper positive charge <strong>of</strong> a typical dipole may become<br />

exposed <strong>to</strong> the ground through differential advection <strong>of</strong><br />

charged hydrometeors <strong>in</strong> the presence <strong>of</strong> strong deeplayer<br />

shear, thus allow<strong>in</strong>g CG lightn<strong>in</strong>g <strong>to</strong> emanate<br />

from anvil level. The <strong>in</strong>terested reader is referred <strong>to</strong><br />

Williams (2001) for a more complete review <strong>of</strong> various<br />

hypotheses that seek <strong>to</strong> expla<strong>in</strong> how s<strong>to</strong>rm dynamics<br />

and microphysics might control CG lightn<strong>in</strong>g polarity,<br />

particularly <strong>in</strong> positive s<strong>to</strong>rms. More recent discussion<br />

and hypotheses regard<strong>in</strong>g the control <strong>of</strong> CG lightn<strong>in</strong>g<br />

polarity <strong>in</strong> severe s<strong>to</strong>rms can be found <strong>in</strong> GW02, Carey<br />

et al. (2003a), CRP03, Knupp et al. (2003), Wiens et al.<br />

(2005), and W05, among others.<br />

2 In ord<strong>in</strong>ary (<strong>in</strong>verted) polarity s<strong>to</strong>rms, the highest charge<br />

layer, which is likely a screen<strong>in</strong>g layer, is negative (positive) as<br />

shown by S<strong>to</strong>lzenburg et al. (1998) (Rust and MacGorman 2002).<br />

b. The meteorological environment and<br />

cloud-<strong>to</strong>-ground lightn<strong>in</strong>g polarity<br />

Given various hypotheses connect<strong>in</strong>g cloud electrification<br />

<strong>to</strong> microphysical and dynamical processes, a<br />

handful <strong>of</strong> studies have explored the detailed relationship<br />

between the local meteorological environment and<br />

the CG lightn<strong>in</strong>g polarity <strong>of</strong> severe s<strong>to</strong>rms (Reap and<br />

MacGorman 1989; Curran and Rust 1992; Smith et al.<br />

2000; GW02). From a comparison <strong>of</strong> numerical model<br />

output with radar and lightn<strong>in</strong>g data, Reap and<br />

MacGorman (1989) found that freez<strong>in</strong>g level height<br />

and deep-layer w<strong>in</strong>d shear were not as important as<br />

boundary layer fields such as moisture convergence,<br />

cyclonic relative vorticity, and strong upward vertical<br />

motions <strong>in</strong> determ<strong>in</strong><strong>in</strong>g the location <strong>of</strong> positive lightn<strong>in</strong>g<br />

activity. Us<strong>in</strong>g multiple proximity sound<strong>in</strong>gs for a<br />

s<strong>in</strong>gle CG lightn<strong>in</strong>g-produc<strong>in</strong>g <strong>to</strong>rnadic s<strong>to</strong>rm, Curran<br />

and Rust (1992) found that strong deep-layer w<strong>in</strong>d<br />

shear may be a necessary, but not sufficient, condition<br />

for the production <strong>of</strong> positive ground flashes. In a study<br />

<strong>of</strong> three <strong>to</strong>rnadic outbreaks, Smith et al. (2000) used<br />

hourly surface airways observation data <strong>to</strong> show that<br />

positive (negative) s<strong>to</strong>rms occurred <strong>in</strong> a strong (weak)<br />

gradient region <strong>of</strong> the surface equivalent potential temperature<br />

( e ), upstream (over or downstream) <strong>of</strong> a surface<br />

e maximum or ridge. A transition from positive <strong>to</strong><br />

negative polarity CG lightn<strong>in</strong>g occurred when <strong>in</strong>itially<br />

positive s<strong>to</strong>rms crossed the center <strong>of</strong> the e ridge <strong>in</strong><strong>to</strong><br />

decreas<strong>in</strong>g values <strong>of</strong> e . Assum<strong>in</strong>g surface e is well<br />

correlated <strong>to</strong> convective available potential energy<br />

(CAPE; e.g., Williams and Renno 1993) and hence <strong>to</strong><br />

the potential maximum updraft speed via parcel theory,<br />

Smith et al. (2000) suggest that dom<strong>in</strong>ant CG lightn<strong>in</strong>g<br />

polarity may be associated with rapid updraft <strong>in</strong>tensification<br />

brought about by an <strong>in</strong>crease <strong>in</strong> the buoyancy<br />

<strong>of</strong> low-level <strong>in</strong>flow air. Inversely, dom<strong>in</strong>ant CG<br />

polarity may be associated with rapid updraft weaken<strong>in</strong>g<br />

brought about by precipitation load<strong>in</strong>g and a decrease<br />

<strong>in</strong> <strong>in</strong>flow buoyancy. In a study <strong>of</strong> a <strong>to</strong>rnado outbreak<br />

us<strong>in</strong>g mobile mesonet and proximity sound<strong>in</strong>gs,<br />

GW02 found that boundary-cross<strong>in</strong>g supercells transitioned<br />

from dom<strong>in</strong>ant CG <strong>to</strong> dom<strong>in</strong>ant CG lightn<strong>in</strong>g<br />

when the s<strong>to</strong>rms experienced enhanced CAPE<br />

(below the <strong>in</strong>-cloud freez<strong>in</strong>g level), boundary layer mix<strong>in</strong>g<br />

ratios, and low-level (0–3 km) vertical w<strong>in</strong>d shear<br />

on the immediate cool side <strong>of</strong> the boundary. Larger<br />

CAPE and 0–3-km shear <strong>in</strong> positive s<strong>to</strong>rms would result<br />

<strong>in</strong> both stronger thermodynamic and dynamic forc<strong>in</strong>g<br />

<strong>of</strong> the updraft and more <strong>in</strong>tense updrafts <strong>in</strong> the<br />

mixed-phase zone (e.g., Weisman and Klemp 1982,<br />

1984, 1986; McCaul and Weisman 2001). Rotunno et al.<br />

(1988) present a theory that squall-l<strong>in</strong>e strength and


1330 MONTHLY WEATHER REVIEW VOLUME 135<br />

longevity were most sensitive <strong>to</strong> the magnitude <strong>of</strong> the<br />

low-level (0–3 km AGL) vertical w<strong>in</strong>d shear perpendicular<br />

<strong>to</strong> squall-l<strong>in</strong>e orientation (i.e., the RKW<br />

theory). An ‘‘optimal’’ state was proposed by RKW<br />

based on the relative strength <strong>of</strong> the circulation associated<br />

with the s<strong>to</strong>rm-generated cold pool and low-level<br />

shear. The deepest lift<strong>in</strong>g and most effective convective<br />

retrigger<strong>in</strong>g occurred when these circulations were <strong>in</strong><br />

near balance. RKW theory was recently reexam<strong>in</strong>ed<br />

and supported by additional 2D and 3D model simulations<br />

<strong>of</strong> Weisman and Rotunno (2004). RKW theory<br />

may expla<strong>in</strong> why 0–3-km shear could affect the <strong>in</strong>tensity<br />

and longevity <strong>of</strong> squall-l<strong>in</strong>e (multicell) convection<br />

and hence cloud electrification and lightn<strong>in</strong>g.<br />

A few studies have drawn <strong>in</strong>ferences regard<strong>in</strong>g the<br />

relationship between CG lightn<strong>in</strong>g polarity and the meteorological<br />

environment over the CONUS us<strong>in</strong>g largescale<br />

clima<strong>to</strong>logical data (Knapp 1994; CRP03; W05).<br />

Knapp (1994) noticed that regions <strong>in</strong> the CONUS<br />

where the troposphere tends <strong>to</strong> be relatively moist (e.g.,<br />

east, southeast, southern pla<strong>in</strong>s) experienced significantly<br />

fewer positive s<strong>to</strong>rms. Us<strong>in</strong>g 10 yr <strong>of</strong> CONUS<br />

National <strong>Lightn<strong>in</strong>g</strong> Detection Network (NLDN) and<br />

the National Centers for <strong>Environmental</strong> Prediction<br />

(NCEP) reanalysis data, CRP03 showed that locations<br />

<strong>of</strong> the monthly frequency maxima <strong>of</strong> severe s<strong>to</strong>rms that<br />

produced predom<strong>in</strong>antly CG and CG lightn<strong>in</strong>g<br />

were systematically <strong>of</strong>fset with respect <strong>to</strong> the clima<strong>to</strong>logical<br />

monthly position <strong>of</strong> the surface e ridge on severe<br />

outbreak days. Positive s<strong>to</strong>rms generally occurred<br />

west and northwest <strong>of</strong> the e ridge <strong>in</strong> the upstream e<br />

gradient region. Negative s<strong>to</strong>rms were primarily located<br />

southeast <strong>of</strong> the positive s<strong>to</strong>rm maximum, closer<br />

<strong>to</strong> the axis <strong>of</strong> the e ridge <strong>in</strong> higher mean values <strong>of</strong> e .<br />

The northward migration <strong>of</strong> this pattern from spr<strong>in</strong>g <strong>to</strong><br />

summer months results <strong>in</strong> the well-known southwest<strong>to</strong>-northeast<br />

(from the Texas panhandle <strong>to</strong> M<strong>in</strong>nesota)<br />

anomaly <strong>in</strong> the percentage <strong>of</strong> CG lightn<strong>in</strong>g <strong>in</strong> annual<br />

maps (e.g., Orville and Huff<strong>in</strong>es 2001) <strong>in</strong> addition <strong>to</strong> the<br />

geographic distribution <strong>of</strong> positive s<strong>to</strong>rms shown <strong>in</strong> Fig.<br />

1 (CRP03; Carey and Rutledge 2003). Us<strong>in</strong>g clima<strong>to</strong>logical<br />

wet bulb potential temperature ( w ) (as a proxy<br />

for CAPE) and cloud-base height (CBH) data, W05<br />

suggest an important role for CBH <strong>in</strong> controll<strong>in</strong>g the<br />

efficiency <strong>of</strong> the transfer <strong>of</strong> CAPE <strong>to</strong> updraft k<strong>in</strong>etic<br />

energy <strong>in</strong> thunders<strong>to</strong>rms. Accord<strong>in</strong>g <strong>to</strong> W05 “an elevated<br />

CBH may enable larger cloud water concentrations<br />

<strong>in</strong> the mixed-phase region, a favorable condition<br />

for the positive charg<strong>in</strong>g <strong>of</strong> large ice particles that may<br />

result <strong>in</strong> thunderclouds with a reversed polarity <strong>of</strong> the<br />

ma<strong>in</strong> cloud dipole.” This condition is also conducive <strong>to</strong><br />

very large hail production (e.g., Knight and Knight<br />

2001). W05 note the geographical co<strong>in</strong>cidence <strong>of</strong> the<br />

clima<strong>to</strong>logical positive s<strong>to</strong>rm, very large hail, dry l<strong>in</strong>e,<br />

and low precipitation (LP) supercell signals over the<br />

central United States. Hence, W05 state that “the comb<strong>in</strong>ed<br />

requirements <strong>of</strong> [large] <strong>in</strong>stability and [high]<br />

CBH serve <strong>to</strong> conf<strong>in</strong>e the region <strong>of</strong> [positive severe<br />

s<strong>to</strong>rms] <strong>to</strong> the vic<strong>in</strong>ity <strong>of</strong> the ridge <strong>in</strong> moist entropy <strong>in</strong><br />

the western Great Pla<strong>in</strong>s.” It is important <strong>to</strong> po<strong>in</strong>t out<br />

that CRP03 found only a weak positive correlation between<br />

hail size and CG fraction. In their study, hail<br />

size expla<strong>in</strong>ed only a m<strong>in</strong>or amount <strong>of</strong> the variance <strong>in</strong><br />

the CG fraction over the United States. Regional<br />

variability <strong>of</strong> CG fraction, regardless <strong>of</strong> hail size, was<br />

dom<strong>in</strong>ant. Thus, one should conclude that the physical<br />

and hence environmental conditions associated with<br />

large hail and CG lightn<strong>in</strong>g production, while potentially<br />

similar, are not uniquely related except <strong>in</strong> specific<br />

circumstances that have yet <strong>to</strong> be determ<strong>in</strong>ed.<br />

There are two underly<strong>in</strong>g arguments associated with<br />

the role <strong>of</strong> CBH <strong>in</strong> the W05 hypothesis—one dynamical<br />

and one microphysical. Fundamentally important <strong>to</strong><br />

both arguments is the observation that water load<strong>in</strong>g<br />

and entra<strong>in</strong>ment can significantly reduce the updraft<br />

velocity below what would be realized by CAPE from<br />

parcel theory, as was shown by Jorgensen and LeMone<br />

(1989) for tropical oceanic convection. W05 suggest a<br />

potential role for CBH <strong>in</strong> modulat<strong>in</strong>g water load<strong>in</strong>g and<br />

entra<strong>in</strong>ment and hence updraft strength. Others (e.g.,<br />

Lucas et al. 1994, 1996; Michaud 1996, 1998; Williams<br />

and Stanfill 2002; Zipser 2003) have made similar suggestions.<br />

In the broad hypothesis, the diameter <strong>of</strong> the<br />

convective updraft is assumed <strong>to</strong> scale with the CBH or<br />

cumulus-<strong>to</strong>pped boundary layer depth (e.g., Kaimal et<br />

al. 1976). Higher CBH is hypothesized <strong>to</strong> result <strong>in</strong> a<br />

larger updraft diameter, less entra<strong>in</strong>ment (e.g., McCarthy<br />

1974), more efficient process<strong>in</strong>g <strong>of</strong> CAPE, stronger<br />

updrafts, and ultimately larger liquid water content<br />

(LWC) <strong>in</strong> the mixed-phase zone. It is <strong>in</strong>terest<strong>in</strong>g <strong>to</strong> note<br />

that Lang and Rutledge (2002) determ<strong>in</strong>ed that a<br />

broad, strong updraft was a necessary <strong>in</strong>gredient for the<br />

production <strong>of</strong> enhanced CG lightn<strong>in</strong>g. A higher CBH<br />

also implies a lesser warm cloud depth (WCD). A<br />

smaller WCD reduces the depth through which the<br />

warm ra<strong>in</strong> (i.e., collision–coalescence) process can occur<br />

<strong>in</strong> the updraft, leav<strong>in</strong>g a greater relative portion <strong>of</strong><br />

cloud liquid water or greater (cloud water)/(ra<strong>in</strong>water)<br />

ratio. All else be<strong>in</strong>g equal, a higher CBH should result<br />

<strong>in</strong> larger supercooled liquid water content <strong>in</strong> the mixedphase<br />

and charg<strong>in</strong>g zone (0° <strong>to</strong> 40°C) because <strong>of</strong> subsequent<br />

ra<strong>in</strong>out or freez<strong>in</strong>g <strong>of</strong> large ra<strong>in</strong> drops (Rosenfeld<br />

and Woodley 2003).<br />

Us<strong>in</strong>g idealized sound<strong>in</strong>gs and numerical cloud simu-


APRIL 2007 C A R E Y A N D B U F F A L O 1331<br />

lations, McCaul and Cohen (2002) demonstrated that<br />

when the level <strong>of</strong> free convection (LFC), which is equal<br />

<strong>to</strong> the lift<strong>in</strong>g condensation level (LCL) <strong>in</strong> much <strong>of</strong> their<br />

study, is with<strong>in</strong> an optimal range <strong>of</strong> 1.5 <strong>to</strong> 2.5 km above<br />

ground level (AGL), maximum convective updraft<br />

overturn<strong>in</strong>g efficiency can approach 100% <strong>of</strong> parcel<br />

theory, while for lower LFC LCL heights the overturn<strong>in</strong>g<br />

efficiency is reduced significantly. The <strong>in</strong>creased<br />

overturn<strong>in</strong>g efficiency with LFC height was associated<br />

with a larger diameter updraft and mean e <strong>of</strong><br />

the low-level updraft. In a sensitivity study, McCaul and<br />

Cohen (2002) found that the strongest updraft case was<br />

associated with a high LFC and low LCL scenario.<br />

From their study, one would expect the most efficient<br />

process<strong>in</strong>g <strong>of</strong> CAPE, strongest updrafts, and least dilution<br />

<strong>of</strong> cloud water from high LFC and especially high<br />

LFC/low LCL environments. The <strong>in</strong>verted-dipole hypothesis<br />

discussed above would thus predict that high<br />

LFC (and low LCL) would be associated with positive<br />

s<strong>to</strong>rms, all else be<strong>in</strong>g equal. In a related study, McCaul<br />

et al. (2005) demonstrated that stronger peak updraft<br />

speeds occurred <strong>in</strong> lower precipitable water (PW) environments,<br />

all else be<strong>in</strong>g equal. The stronger peak updrafts<br />

were apparently the result <strong>of</strong> less ra<strong>in</strong>water load<strong>in</strong>g<br />

and the lower altitudes at which the latent heat <strong>of</strong><br />

freez<strong>in</strong>g and deposition commences <strong>in</strong> relatively lower<br />

PW environments. Hence, PW could possibly modulate<br />

updraft strength <strong>in</strong>dependent <strong>of</strong> CBH.<br />

S<strong>in</strong>ce it is difficult <strong>to</strong> obta<strong>in</strong> representative convective<br />

<strong>in</strong>flow sound<strong>in</strong>gs, only a few studies have analyzed<br />

representative thermodynamic and dynamic characteristics<br />

<strong>of</strong> the mesoscale environment <strong>in</strong> conjunction with<br />

CG lightn<strong>in</strong>g properties. Clearly, more field observations<br />

and further study is warranted <strong>in</strong> order <strong>to</strong> test and,<br />

if necessary, ref<strong>in</strong>e the above hypotheses l<strong>in</strong>k<strong>in</strong>g environmental<br />

conditions <strong>to</strong> updraft <strong>in</strong>tensity, supercooled<br />

cloud water contents, electrification, and ultimately CG<br />

lightn<strong>in</strong>g polarity. Therefore, this study seeks <strong>to</strong> <strong>in</strong>vestigate<br />

the relationship between CG lightn<strong>in</strong>g polarity<br />

and the immediate meteorological environment <strong>of</strong> severe<br />

s<strong>to</strong>rms, thereby provid<strong>in</strong>g further <strong>in</strong>sight <strong>in</strong><strong>to</strong> why<br />

only some severe s<strong>to</strong>rms are dom<strong>in</strong>ated by CG<br />

flashes, and <strong>in</strong> particular, what conditions lead <strong>to</strong> this<br />

CG dom<strong>in</strong>ance. A determ<strong>in</strong>ation <strong>of</strong> whether environmental<br />

conditions are systematically related <strong>to</strong> CG<br />

lightn<strong>in</strong>g polarity, and if so, what these conditions are,<br />

is a crucial step <strong>in</strong> determ<strong>in</strong><strong>in</strong>g the reliability <strong>of</strong> us<strong>in</strong>g<br />

NLDN real-time flash polarity data for the short-term<br />

prediction <strong>of</strong> severe weather. Furthermore, determ<strong>in</strong><strong>in</strong>g<br />

the relationship between certa<strong>in</strong> environmental<br />

conditions and CG lightn<strong>in</strong>g polarity will lead <strong>to</strong> an<br />

improved understand<strong>in</strong>g <strong>of</strong> cloud electrification mechanisms,<br />

especially <strong>in</strong> positive s<strong>to</strong>rms, which rema<strong>in</strong>s<br />

speculative at this time (Williams 2001).<br />

3. Data and methodology<br />

Us<strong>in</strong>g data from the International H 2 0 Project<br />

(IHOP_2002; Weckwerth et al. 2004), we explored the<br />

relationship between the local meteorological environment<br />

and the CG lightn<strong>in</strong>g polarity <strong>of</strong> severe s<strong>to</strong>rms. In<br />

particular, we wished <strong>to</strong> determ<strong>in</strong>e if there were significant<br />

and systematic differences <strong>in</strong> the mesoscale environments<br />

<strong>of</strong> negative and positive severe s<strong>to</strong>rms that<br />

are consistent with the hypothesis discussed above. In a<br />

more general sense, our objective was <strong>to</strong> test if there<br />

was a significant relationship between CG lightn<strong>in</strong>g polarity<br />

and meteorological parameters that are known <strong>to</strong><br />

affect s<strong>to</strong>rm dynamics and microphysics, and hence<br />

possibly cloud electrification and lightn<strong>in</strong>g. IHOP_2002<br />

was conducted from 13 May <strong>to</strong> 25 June 2002 across the<br />

Southern Great Pla<strong>in</strong>s (Kansas, Oklahoma, and the<br />

Texas panhandle—see boxed area <strong>in</strong> Fig. 1). The<br />

IHOP_2002 area also happens <strong>to</strong> be ideal for our scientific<br />

objectives because it represents a transition region<br />

between the prevalence <strong>of</strong> positive and negative<br />

severe s<strong>to</strong>rms <strong>in</strong> the central CONUS (see elliptical area<br />

<strong>in</strong> Fig. 1), allow<strong>in</strong>g us <strong>to</strong> carefully quantify the meteorological<br />

environment <strong>of</strong> both negative and positive<br />

s<strong>to</strong>rms. The ma<strong>in</strong> goal <strong>of</strong> IHOP_2002 was <strong>to</strong> obta<strong>in</strong><br />

more accurate and reliable measurements <strong>of</strong> moisture<br />

<strong>in</strong> the air, <strong>in</strong> an attempt <strong>to</strong> improve quantitative precipitation<br />

forecasts and <strong>in</strong>crease understand<strong>in</strong>g <strong>of</strong> convective<br />

<strong>in</strong>itiation (Weckwerth et al. 2004). Thus, detailed<br />

measurements <strong>of</strong> the mesoscale environment<br />

(both w<strong>in</strong>d and thermodynamic parameters) <strong>in</strong> both the<br />

horizontal and vertical were obta<strong>in</strong>ed dur<strong>in</strong>g<br />

IHOP_2002. Of particular <strong>in</strong>terest <strong>to</strong> this study is the<br />

multitude <strong>of</strong> environmental sound<strong>in</strong>gs taken dur<strong>in</strong>g<br />

IHOP_2002.<br />

The relationship between CG lightn<strong>in</strong>g polarity and<br />

the local mesoscale environment was <strong>in</strong>vestigated for 6<br />

different days (23, 24 May 2002; 4, 12, 15, 19 June 2002)<br />

dur<strong>in</strong>g IHOP_2002. The chosen days were selected<br />

based on the prevalence <strong>of</strong> positive and/or negative<br />

s<strong>to</strong>rms (Table 1), the availability <strong>of</strong> proximity sound<strong>in</strong>g<br />

data <strong>to</strong> characterize the local mesoscale environments<br />

<strong>of</strong> these s<strong>to</strong>rms (Table 2), and the occurrence <strong>of</strong> severe<br />

weather (Table 3). Dur<strong>in</strong>g these 6 days, n<strong>in</strong>e dist<strong>in</strong>ct<br />

mesoscale regions were identified based on the CG<br />

lightn<strong>in</strong>g polarity characteristics <strong>of</strong> the s<strong>to</strong>rms occurr<strong>in</strong>g<br />

with<strong>in</strong> them. Four <strong>of</strong> these regions conta<strong>in</strong>ed positive<br />

s<strong>to</strong>rms and the rema<strong>in</strong><strong>in</strong>g five were associated with<br />

negative s<strong>to</strong>rms as def<strong>in</strong>ed earlier (Table 1). All <strong>of</strong> the<br />

regions conta<strong>in</strong>ed significant numbers <strong>of</strong> severe s<strong>to</strong>rm


1332 MONTHLY WEATHER REVIEW VOLUME 135<br />

TABLE 1. Def<strong>in</strong>ition <strong>of</strong> mesoscale regions with<strong>in</strong> the IHOP_2002 doma<strong>in</strong> and characterization <strong>of</strong> CG lightn<strong>in</strong>g. The overall<br />

percentage <strong>of</strong> CG lightn<strong>in</strong>g, mean <strong>to</strong>tal CG lightn<strong>in</strong>g flash density, and number <strong>of</strong> positive and negative CG flashes are given.<br />

Positive mesoscale regions—Mesoscale regions conta<strong>in</strong><strong>in</strong>g 25% CG lightn<strong>in</strong>g<br />

Date Time (UTC) Lat–Lon CG%<br />

Mean flash density<br />

(flashes km 2 yr 1 ) Positive flashes Negative flashes<br />

23 May 1800–0300 33°–38°N, 103° <strong>to</strong> 100°W 60.7 0.146 722 467<br />

24 May 2000–0400 33.5°–37°N, 101.5° <strong>to</strong> 98.5°W 32.2 0.183 1783 3755<br />

15 Jun 1800–0300 33°–39°N, 103° <strong>to</strong> 99°W 43.4 0.099 3395 4436<br />

19 Jun 1800–0300 37°–43°N, 103° <strong>to</strong> 97°W 71.5 0.160 4540 1807<br />

Negative mesoscale regions—Mesoscale regions conta<strong>in</strong><strong>in</strong>g 25% CG lightn<strong>in</strong>g<br />

Date Time (UTC) Lat–Lon CG%<br />

Mean flash density<br />

(flashes km 2 yr 1 ) Positive flashes Negative flashes<br />

23 May 1800–0300 34°–40°N, 100° <strong>to</strong> 94°W 6.5 0.126 540 7760<br />

24 May 2000–0400 32.5°–38°N, 98.5° <strong>to</strong> 95°W 7.5 0.186 790 9757<br />

4 Jun 1200–0100 33°–40°N, 103° <strong>to</strong> 95°W 9.2 0.210 3414 33 819<br />

12 Jun 2000–0400 32°–39°N, 103° <strong>to</strong> 95°W 8.9 0.384 3340 34 169<br />

15 Jun 1800–0300 33°–39°N, 99° <strong>to</strong> 95.5°W 17.1 0.104 1418 6865<br />

reports except for the negative mesoscale region on 23<br />

May 2002, which conta<strong>in</strong>ed only one large hail report<br />

(Table 3).<br />

a. <strong>Cloud</strong>-<strong>to</strong>-ground lightn<strong>in</strong>g data<br />

<strong>Cloud</strong>-<strong>to</strong>-ground lightn<strong>in</strong>g data analyzed <strong>in</strong> this study<br />

were collected by the NLDN (Cumm<strong>in</strong>s et al. 1998),<br />

which is owned and operated by Vaisala (Tucson, Arizona).<br />

The NLDN records the time, location, polarity,<br />

peak current, and multiplicity (number <strong>of</strong> strokes per<br />

flash) <strong>of</strong> CG lightn<strong>in</strong>g flashes. The time, location, polarity,<br />

and peak current reported for a flash are those<br />

measured for the first return stroke <strong>of</strong> the flash. Over<br />

most <strong>of</strong> the cont<strong>in</strong>ental United States, <strong>in</strong>clud<strong>in</strong>g the<br />

IHOP_2002 doma<strong>in</strong> (Fig. 1), the NLDN has a median<br />

location accuracy <strong>of</strong> 0.5 km and a flash detection efficiency<br />

<strong>of</strong> 80%–90% for strokes with peak currents<br />

greater than 5 kA (Cumm<strong>in</strong>s et al. 1998). S<strong>in</strong>ce CG<br />

flashes with peak currents less than 10 kA were likely<br />

associated with misidentified cloud flashes (Cumm<strong>in</strong>s<br />

TABLE 2. Total number <strong>of</strong> sound<strong>in</strong>gs used <strong>to</strong> characterize each mesoscale region, and the number <strong>of</strong> these sound<strong>in</strong>gs that were full<br />

and truncated. The pressure level <strong>of</strong> truncation (hPa) is also provided, with the number <strong>in</strong> parentheses follow<strong>in</strong>g the pressure level<br />

<strong>in</strong>dicat<strong>in</strong>g the number <strong>of</strong> sound<strong>in</strong>gs truncated at that respective level (unless only one sound<strong>in</strong>g was truncated at the given level).<br />

Sound<strong>in</strong>g numbers and levels <strong>of</strong> truncation are also provided for the <strong>to</strong>tal number <strong>of</strong> positive (negative) sound<strong>in</strong>gs from all positive<br />

(negative) mesoscale regions comb<strong>in</strong>ed.<br />

Date<br />

Positive mesoscale regions—Mesoscale regions conta<strong>in</strong><strong>in</strong>g 25% CG lightn<strong>in</strong>g<br />

Tot No.<br />

<strong>of</strong> sound<strong>in</strong>gs<br />

Full<br />

sound<strong>in</strong>gs<br />

Truncated<br />

sound<strong>in</strong>gs<br />

Truncation level (hPa)<br />

23 May 1 1 0 N/A<br />

24 May 8 1 7 350, 505, 575 (5)<br />

15 Jun 7 4 3 340, 440, 485<br />

19 Jun 8 4 4 365, 455 (3)<br />

All positive sound<strong>in</strong>gs 24 10 14 340, 350, 365, 440, 455 (3), 485, 505, 575 (5)<br />

Date<br />

Negative mesoscale regions—Mesoscale regions conta<strong>in</strong><strong>in</strong>g 25% CG lightn<strong>in</strong>g<br />

Tot No.<br />

<strong>of</strong> sound<strong>in</strong>gs<br />

Full<br />

sound<strong>in</strong>gs<br />

Truncated<br />

sound<strong>in</strong>gs<br />

Truncation level (hPa)<br />

23 May 3 3 0 N/A<br />

24 May 3 3 0 N/A<br />

4 Jun 9 9 0 N/A<br />

12 Jun 8 6 2 450, 485<br />

15 Jun 1 1 0 N/A<br />

All negative sound<strong>in</strong>gs 24 22 2 450, 485


APRIL 2007 C A R E Y A N D B U F F A L O 1333<br />

TABLE 3. Description <strong>of</strong> s<strong>to</strong>rm type and severity with<strong>in</strong> each mesoscale region. Number <strong>of</strong> severe w<strong>in</strong>d (26 m s 1 ), hail (diameter<br />

1.9 cm), and <strong>to</strong>rnado (F0–F5) reports dur<strong>in</strong>g the analysis time period listed <strong>in</strong> Table 1 for each mesoscale region are given.<br />

Positive mesoscale regions—Mesoscale regions conta<strong>in</strong><strong>in</strong>g 25% CG lightn<strong>in</strong>g<br />

Date S<strong>to</strong>rm type(s) Severe W<strong>in</strong>d Hail Tornado<br />

23 May L<strong>in</strong>e <strong>of</strong> supercells Yes 1 37 3 (all F0)<br />

24 May Squall l<strong>in</strong>e (ord<strong>in</strong>ary cells with several embedded supercells) Yes 3 49 2 (both F0)<br />

15 Jun Multicell (ord<strong>in</strong>ary and one supercell) evolv<strong>in</strong>g <strong>in</strong><strong>to</strong> squall l<strong>in</strong>e Yes 60 60 3 (all F0)<br />

19 Jun Broken squall l<strong>in</strong>e <strong>of</strong> ord<strong>in</strong>ary cells; isolated supercell Yes 21 31 9 (6 F0, 3 F1)<br />

Negative mesoscale regions—Mesoscale regions conta<strong>in</strong><strong>in</strong>g 25% CG lightn<strong>in</strong>g<br />

Date S<strong>to</strong>rm type(s) Severe W<strong>in</strong>d Hail Tornado<br />

23 May Broad cluster <strong>of</strong> ord<strong>in</strong>ary multicell convection No* 0 1 0<br />

24 May Broken squall l<strong>in</strong>e <strong>of</strong> ord<strong>in</strong>ary cells Yes 4 21 0<br />

4 Jun Squall l<strong>in</strong>e (ord<strong>in</strong>ary cells with two supercells) evolv<strong>in</strong>g <strong>in</strong><strong>to</strong> MCS Yes 22 70 1 (F0)<br />

12 Jun Scattered supercells evolv<strong>in</strong>g <strong>in</strong><strong>to</strong> squall l<strong>in</strong>e Yes 40 55 6 (5 F0, 1 F1)<br />

15 Jun Multicell (ord<strong>in</strong>ary and one supercell) evolv<strong>in</strong>g <strong>in</strong><strong>to</strong> squall l<strong>in</strong>e Yes 66 28 1 (F0)<br />

* Despite the one severe hail report with<strong>in</strong> the 23 May negative mesoscale region, the region was classified as nonsevere, s<strong>in</strong>ce this one<br />

<strong>in</strong>cidence <strong>of</strong> severe weather was an exception <strong>to</strong> the rest <strong>of</strong> the nonsevere convection <strong>in</strong> the mesoscale region.<br />

et al. 1998; Wacker and Orville 1999a,b), they were<br />

removed from the data sample as suggested by Cumm<strong>in</strong>s<br />

et al. (1998).<br />

Maps <strong>of</strong> the CG percentage over the IHOP_2002<br />

doma<strong>in</strong> (not shown) were visually <strong>in</strong>spected <strong>to</strong> def<strong>in</strong>e<br />

mesoscale regions across which either positive or negative<br />

s<strong>to</strong>rms prevailed. The overall percentage <strong>of</strong> positive<br />

flashes with<strong>in</strong> each <strong>of</strong> these def<strong>in</strong>ed regions was<br />

then calculated <strong>to</strong> confirm that the mesoscale region<br />

was <strong>in</strong>deed characterized by positive s<strong>to</strong>rms (i.e., overall<br />

percent positive flashes for the mesoscale region<br />

25%; termed positive mesoscale regions) or negative<br />

s<strong>to</strong>rms (i.e., overall percent positive flashes for the mesoscale<br />

region 25%; termed negative mesoscale regions).<br />

The maps were also visually <strong>in</strong>spected <strong>to</strong> ensure<br />

that the CG lightn<strong>in</strong>g polarity was consistent across the<br />

entire mesoscale region. Maps <strong>of</strong> flash density were<br />

used as checks <strong>to</strong> <strong>in</strong>sure that the associated s<strong>to</strong>rms generated<br />

a significant number <strong>of</strong> flashes on a consistent<br />

basis across the def<strong>in</strong>ed mesoscale region. Through this<br />

method, four positive (23, 24 May; 15, 19 June) and five<br />

negative (23, 24 May; 4, 12, 15 June) mesoscale regions<br />

were identified (Table 1). As expected from CRP03<br />

and Fig. 1, negative mesoscale regions were found generally<br />

east <strong>of</strong> positive mesoscale regions on days when<br />

both were present. The positive (negative) s<strong>to</strong>rms were<br />

characterized by a mean CG percentage <strong>of</strong> 50% (9%)<br />

and a range <strong>of</strong> 32%–72% (7%–17%). 3 Even though<br />

negative s<strong>to</strong>rms had generally larger CG flash densities<br />

and rates, the mean CG flash density <strong>in</strong> each region was<br />

comparable (i.e., same order <strong>of</strong> magnitude) because the<br />

negative s<strong>to</strong>rm systems were typically much more widespread<br />

over a larger geographic area than positive<br />

s<strong>to</strong>rm systems (Table 1).<br />

b. Sound<strong>in</strong>g data<br />

Data from several different sound<strong>in</strong>g platforms operat<strong>in</strong>g<br />

dur<strong>in</strong>g IHOP_2002 were used <strong>to</strong> characterize<br />

the n<strong>in</strong>e mesoscale regions <strong>of</strong> <strong>in</strong>terest. Sound<strong>in</strong>g platforms<br />

<strong>in</strong>cluded National Weather Service (NWS) upper-air<br />

sites, Atmospheric Radiation Measurement<br />

<strong>Cloud</strong>s and Radiation Testbed (ARM CART) sites, the<br />

National Center for Atmospheric Research/Atmospheric<br />

Technology Division (NCAR/ATD) Integrated<br />

Sound<strong>in</strong>g System (ISS) facility, the National Severe<br />

S<strong>to</strong>rms Labora<strong>to</strong>ry (NSSL) Mobile Cross-cha<strong>in</strong> Loran<br />

Atmospheric Sound<strong>in</strong>g System (MCLASS) facility, and<br />

NCAR/ATD Mobile GPS/Loran Atmospheric Sound<strong>in</strong>g<br />

System (MGLASS) facilities. Dropsondes launched<br />

from a Flight International (FI) Learjet were also used.<br />

Sound<strong>in</strong>g data from these platforms were quality con-<br />

3 Cumm<strong>in</strong>s et al. (2006) reported that the recent NLDN upgrade<br />

<strong>in</strong>creased the detection <strong>of</strong> low-amplitude flashes and thus<br />

the potential for misclassify<strong>in</strong>g cloud flashes as ground discharges<br />

<strong>of</strong> either polarity. To test the sensitivity <strong>of</strong> our regional CG polarity<br />

classification <strong>to</strong> peak current threshold<strong>in</strong>g, we removed all<br />

NLDN-detected CG flashes with peak currents less than 10, 15,<br />

and 20 kA (K. Cumm<strong>in</strong>s 2006, personal communication). The<br />

mean CG percentage <strong>in</strong> () CG polarity regions after remov<strong>in</strong>g<br />

all CG flashes with peak currents less than 10, 15, and 20 kA<br />

was 53%, 66%, and 81% (10%, 9%, and 11%), respectively. S<strong>in</strong>ce<br />

our CG polarity classification was <strong>in</strong>sensitive <strong>to</strong> the choice <strong>of</strong> a<br />

peak current threshold, we have chosen <strong>to</strong> follow Cumm<strong>in</strong>s et al.<br />

(1998) until ongo<strong>in</strong>g NLDN flash classification improvements are<br />

completed.


1334 MONTHLY WEATHER REVIEW VOLUME 135<br />

trolled and <strong>in</strong>terpolated <strong>to</strong> a constant vertical resolution<br />

<strong>of</strong> 5 hPa by the University Corporation for Atmospheric<br />

Research (UCAR) Jo<strong>in</strong>t Office for Science<br />

Support (JOSS). See the appendix for details regard<strong>in</strong>g<br />

the selection <strong>of</strong> proximity sound<strong>in</strong>gs.<br />

The National Centers Advanced Weather Interactive<br />

Process<strong>in</strong>g System (AWIPS) Skew-T Hodograph<br />

Analysis and Research Program (NSHARP; Hart and<br />

Korotky 1991) was used for sound<strong>in</strong>g display and analysis,<br />

<strong>in</strong>clud<strong>in</strong>g the calculations <strong>of</strong> most sound<strong>in</strong>g-derived<br />

parameters. NSHARP <strong>in</strong>cludes a virtual temperature<br />

correction (Doswell and Rasmussen 1994) <strong>in</strong> the calculations<br />

<strong>of</strong> thermodynamic parameters <strong>to</strong> account for the<br />

effects <strong>of</strong> water vapor. A mean-layer parcel, us<strong>in</strong>g mean<br />

temperature and dewpo<strong>in</strong>t <strong>in</strong> the lowest 100 hPa (approximately<br />

1 km <strong>in</strong> depth), was used <strong>to</strong> calculate thermodynamic<br />

parameters. Craven et al. (2002a) determ<strong>in</strong>ed<br />

that a mean-layer parcel is more representative<br />

<strong>of</strong> the actual parcel associated with convective cloud<br />

development than is a surface-based parcel, and thus<br />

recommended us<strong>in</strong>g a mean-layer parcel <strong>in</strong> the calculations<br />

<strong>of</strong> thermodynamic parameters. Thompson et al.<br />

(2003) also found a mean-layer parcel <strong>to</strong> be superior <strong>to</strong><br />

a surface-based parcel <strong>in</strong> accurately calculat<strong>in</strong>g thermodynamic<br />

parameters <strong>in</strong> convective s<strong>to</strong>rm environments.<br />

Equivalent potential temperature ( e ) was computed<br />

from the recommended formula <strong>in</strong> Bol<strong>to</strong>n (1980).<br />

CAPE is traditionally def<strong>in</strong>ed and calculated from the<br />

LFC <strong>to</strong> the equilibrium level (EL; Moncrieff and Miller<br />

1976). For our comparison with CG lightn<strong>in</strong>g, CAPE<br />

for a specific temperature layer was calculated by modify<strong>in</strong>g<br />

the sound<strong>in</strong>gs <strong>in</strong>put <strong>to</strong> NSHARP. For example, <strong>to</strong><br />

calculate “CAPE between the 10°C and 40°C levels,”<br />

we truncated the sound<strong>in</strong>g <strong>in</strong>put <strong>to</strong> NSHARP at<br />

40°C, repeated the process with a sound<strong>in</strong>g truncated<br />

at 10°C, and then subtracted the calculated CAPE<br />

value <strong>of</strong> the latter from the former NSHARP run.<br />

In calculat<strong>in</strong>g environmental parameters from the<br />

sound<strong>in</strong>g data, obviously only those sound<strong>in</strong>gs that conta<strong>in</strong>ed<br />

the necessary data and extended through the<br />

necessary depth <strong>to</strong> accurately calculate each respective<br />

parameter were used. For <strong>in</strong>stance, CAPE (i.e., LFC <strong>to</strong><br />

EL) was not calculated from sound<strong>in</strong>gs truncated below<br />

the EL. As another example, s<strong>to</strong>rm system motion,<br />

which is estimated from the mean w<strong>in</strong>d <strong>in</strong> the surface <strong>to</strong><br />

6 km AGL layer, could not be estimated by NSHARP<br />

for sound<strong>in</strong>gs that were truncated below 6 km AGL. As<br />

a result, any parameters dependent on s<strong>to</strong>rm system<br />

motion (e.g., s<strong>to</strong>rm relative w<strong>in</strong>ds) were also discarded<br />

from such sound<strong>in</strong>gs. Some <strong>of</strong> the mobile (MCLASS,<br />

MGLASS) sound<strong>in</strong>gs launched dur<strong>in</strong>g IHOP_2002, as<br />

well as the dropsondes, did not extend through the full<br />

depth <strong>of</strong> the troposphere, s<strong>in</strong>ce IHOP_2002 <strong>in</strong>vestiga<strong>to</strong>rs<br />

were primarily concerned with measur<strong>in</strong>g low-level<br />

moisture fields. Hence, some IHOP_2002 sound<strong>in</strong>gs<br />

could not be used for calculat<strong>in</strong>g parameters that require<br />

measurements through a significant depth <strong>of</strong> the<br />

troposphere [e.g., <strong>to</strong>tal CAPE, s<strong>to</strong>rm motion and dependent<br />

parameters, deep-layer (0–6 km AGL) shear,<br />

PW, etc.]. Table 2 displays the <strong>to</strong>tal number <strong>of</strong> sound<strong>in</strong>gs<br />

used <strong>to</strong> characterize each mesoscale region, the<br />

number <strong>of</strong> these sound<strong>in</strong>gs that were full (extended<br />

through the depth <strong>of</strong> the troposphere), the number that<br />

were truncated (did not extend through the depth <strong>of</strong><br />

the troposphere), and their truncation levels. As seen <strong>in</strong><br />

Table 2, all mesoscale regions <strong>in</strong>cluded at least one full<br />

sound<strong>in</strong>g so that all environmental parameters could be<br />

calculated for each region. In the positive (negative)<br />

regions, there were an average <strong>of</strong> 6.0 (4.8) <strong>to</strong>tal and 2.5<br />

(4.4) full tropospheric sound<strong>in</strong>gs per region. This is a<br />

rather small number <strong>of</strong> sound<strong>in</strong>gs per region for a statistically<br />

significant comparison <strong>of</strong> environmental conditions<br />

between <strong>in</strong>dividual mesoscale regions. As a result,<br />

we comb<strong>in</strong>ed all 24, <strong>in</strong>clud<strong>in</strong>g 10 full, positive region<br />

sound<strong>in</strong>gs <strong>in</strong><strong>to</strong> one dist<strong>in</strong>ct “positive group” and<br />

all 24, <strong>in</strong>clud<strong>in</strong>g 22 full, negative region sound<strong>in</strong>gs <strong>in</strong><strong>to</strong><br />

another dist<strong>in</strong>ct “negative group” that could then be<br />

compared statistically.<br />

In characteriz<strong>in</strong>g the mesoscale environments <strong>of</strong><br />

positive and negative s<strong>to</strong>rms, special emphasis was<br />

placed on those thermodynamic and w<strong>in</strong>d parameters<br />

that allowed test<strong>in</strong>g <strong>of</strong> the hypothesis discussed earlier.<br />

The hypothesis states that the local mesoscale environment<br />

<strong>in</strong>directly <strong>in</strong>fluences CG lightn<strong>in</strong>g polarity by directly<br />

controll<strong>in</strong>g s<strong>to</strong>rm structure, dynamics, and microphysics,<br />

while the associated corollary more specifically<br />

states that broad, <strong>in</strong>tense updrafts and associated high<br />

LWCs <strong>in</strong> positive s<strong>to</strong>rms lead <strong>to</strong> positive (negative)<br />

charg<strong>in</strong>g <strong>of</strong> graupel and hail (ice crystals) <strong>in</strong> mixedphase<br />

conditions via the NIC mechanism, an <strong>in</strong>vertedpolarity<br />

charge region, and <strong>in</strong>creased frequency <strong>of</strong><br />

CG lightn<strong>in</strong>g. Based on this hypothesis and corollary,<br />

those environmental parameters that strongly <strong>in</strong>fluence<br />

s<strong>to</strong>rm organization, updraft <strong>in</strong>tensity, and associated<br />

cloud LWCs were emphasized. To assess updraft <strong>in</strong>tensity<br />

from buoyancy, several environmental parameters<br />

were <strong>in</strong>vestigated, <strong>in</strong>clud<strong>in</strong>g CAPE, normalized CAPE<br />

(NCAPE), and lapse rates computed through various<br />

layers <strong>of</strong> the atmosphere, especially those relevant <strong>to</strong><br />

cloud electrification (i.e., between 0° and 40°C).<br />

NCAPE is the mean parcel acceleration associated with<br />

buoyancy <strong>in</strong> a layer, assum<strong>in</strong>g parcel theory (Blanchard<br />

1998). By divid<strong>in</strong>g CAPE by the depth over which<br />

buoyancy is <strong>in</strong>tegrated, NCAPE can account for variations<br />

<strong>in</strong> the shape <strong>of</strong> the vertical pr<strong>of</strong>ile <strong>of</strong> CAPE (e.g.,


APRIL 2007 C A R E Y A N D B U F F A L O 1335<br />

TABLE 4. Mean (median) sound<strong>in</strong>g properties <strong>of</strong> grouped negative and positive mesoscale regions. The number <strong>of</strong> sound<strong>in</strong>gs <strong>in</strong> each<br />

negative (N) and positive (N) parameter sample is also shown. For each parameter, the statistical significance level <strong>of</strong> the difference<br />

<strong>in</strong> means accord<strong>in</strong>g <strong>to</strong> the procedure described <strong>in</strong> section 3 was very highly significant (99.9% level, p 0.001).<br />

Very highly significant: 99.9% level<br />

Grouped negative regions<br />

Grouped positive regions<br />

Mean (median) N Mean (median) N<br />

WCD FL LCL 2950 (3090) m 24 1700 (1650) m 24<br />

LCL 1120 (1080) m AGL 24 2080 (2010) m AGL 24<br />

Mean mix<strong>in</strong>g ratio <strong>in</strong> the lowest 100 hPa 14.0 (14.3) g kg 1 24 10.9 (10.8) g kg 1 24<br />

850–500-hPa lapse rate 7.1°C (7.0°C) km 1 24 8.4°C (8.4°C) km 1 18<br />

Wet-bulb zero (WBZ) height 3280 (3310) m AGL 24 2870 (2850) m AGL 24<br />

PW <strong>in</strong> surface <strong>to</strong> 400-hPa layer 3.6 (3.5) cm 22 2.7 (2.6) cm 13<br />

Surface dewpo<strong>in</strong>t (T d ) 18.7 (18.7)°C 24 14.3 (14.3)°C 24<br />

Surface dewpo<strong>in</strong>t depression (T T d ) 7.8 (7.6)°C 24 16.7 (15.8)°C 24<br />

Lucas et al. 1994) and provide a better estimate <strong>of</strong> updraft<br />

velocity (Blanchard 1998). CAPE, low-level (0–3<br />

km) and deep-layer (0–6 km) vertical w<strong>in</strong>d shear, and<br />

the bulk Richardson number (BRN) were <strong>in</strong>vestigated<br />

with regard <strong>to</strong> s<strong>to</strong>rm organization (e.g., Weisman and<br />

Klemp 1982, 1984, 1986) and dynamic forc<strong>in</strong>g <strong>of</strong> the<br />

updraft (e.g., Weisman and Klemp 1982; McCaul and<br />

Weisman 2001). Depth <strong>of</strong> the warm cloud layer can<br />

affect the amount <strong>of</strong> cloud liquid water available <strong>to</strong> the<br />

mixed-phase region, and hence is hypothesized <strong>to</strong> affect<br />

s<strong>to</strong>rm electrification and CG lightn<strong>in</strong>g polarity as discussed<br />

earlier. Warm cloud layer depth was calculated<br />

by subtract<strong>in</strong>g the LCL height (a measure <strong>of</strong> CBH)<br />

from the height <strong>of</strong> the 0°C level. (A complete list <strong>of</strong><br />

calculated sound<strong>in</strong>g parameters can be found <strong>in</strong> Tables<br />

4–8). Formally def<strong>in</strong><strong>in</strong>g all <strong>of</strong> these environmental variables<br />

and describ<strong>in</strong>g their impact on the dynamics and<br />

microphysics <strong>of</strong> convection is beyond the scope <strong>of</strong> this<br />

paper. The <strong>in</strong>terested reader is referred <strong>to</strong> Weisman<br />

and Klemp (1986), Houze (1993), and <strong>to</strong> more recent<br />

papers relat<strong>in</strong>g sound<strong>in</strong>g parameters <strong>to</strong> severe convective<br />

weather (e.g., Rasmussen and Blanchard 1998;<br />

Moller 2001; Rasmussen 2003) for an overview.<br />

The samples associated with positive and negative<br />

s<strong>to</strong>rm environments for each parameter were explored<br />

us<strong>in</strong>g standard statistical techniques (e.g., Wilks 1995).<br />

In particular, the samples for each parameter <strong>in</strong> each<br />

type <strong>of</strong> environment were compared for location us<strong>in</strong>g<br />

the arithmetic mean and median. Significance test<strong>in</strong>g <strong>of</strong><br />

the difference <strong>in</strong> location was accomplished us<strong>in</strong>g a rigorous<br />

approach. The two samples (i.e., associated with<br />

positive and negative s<strong>to</strong>rm environments) were first<br />

checked for normality us<strong>in</strong>g the Anderson–Darl<strong>in</strong>g test<br />

(Stephens 1974) at the 95% significance level. If both<br />

samples for a given parameter were found <strong>to</strong> come<br />

from a normal distribution, then a two-sample heteroscedastic<br />

(i.e., assum<strong>in</strong>g unequal variances) twotailed<br />

Student’s t test was performed <strong>to</strong> test the null<br />

hypothesis that the two population distribution functions<br />

correspond<strong>in</strong>g <strong>to</strong> the two random samples are<br />

identical aga<strong>in</strong>st the alternative hypothesis that they<br />

differ by location (i.e., their mean and medians are different).<br />

If one or both <strong>of</strong> the samples for a given parameter<br />

were found <strong>to</strong> be nonnormal, then an attempt<br />

was made <strong>to</strong> normalize them us<strong>in</strong>g a suitable Box–Cox<br />

(or power) transformation (Wilks 1995). If the samples<br />

could be reexpressed <strong>to</strong> normality, as judged by an<br />

Anderson–Darl<strong>in</strong>g normality test at the 95% significance<br />

level, then a two-sample heteroscedastic twotailed<br />

Student’s t test was performed on the transformed<br />

data. Otherwise, the nonparametric Wilcoxon–<br />

Mann–Whitney rank sum test (Wilks 1995) was applied<br />

<strong>to</strong> the nonnormal samples <strong>to</strong> test the null hypothesis<br />

that the two population distribution functions correspond<strong>in</strong>g<br />

<strong>to</strong> the two random samples are identical<br />

aga<strong>in</strong>st the alternative hypothesis that they differ by<br />

location. To conduct these analyses, we used a comb<strong>in</strong>ation<br />

<strong>of</strong> Micros<strong>of</strong>t Excel, the ITT Corporation Interactive<br />

Data Language (IDL), and the National Institute<br />

<strong>of</strong> Standards and Technology (NIST) Dataplot (Heckert<br />

and Filliben 2003) data visualization and statistical<br />

analysis packages.<br />

4. Results<br />

a. Overall mean environmental conditions<br />

The overall mean (and median) environmental parameters<br />

for negative and positive mesoscale regions<br />

are presented <strong>in</strong> Tables 4–8, as ranked by the highest<br />

significance level achieved with either the Student’s t<br />

test or the Wilcoxon–Mann–Whitney rank sum test as<br />

described <strong>in</strong> section 3. The mean (median) values listed<br />

for each parameter represent the parameter mean (median)<br />

for all sound<strong>in</strong>gs characteriz<strong>in</strong>g negative s<strong>to</strong>rm<br />

environments (i.e., all five negative mesoscale regions<br />

grouped <strong>to</strong>gether) and all sound<strong>in</strong>gs characteriz<strong>in</strong>g<br />

positive s<strong>to</strong>rm environments (i.e., all four positive mesoscale<br />

regions grouped <strong>to</strong>gether).


1336 MONTHLY WEATHER REVIEW VOLUME 135<br />

TABLE 5. Same as <strong>in</strong> Table 4, but the difference <strong>in</strong> each mean (median) parameter was highly significant<br />

(99% level, 0.001 p 0.01).<br />

Highly significant: 99% level<br />

Grouped negative regions<br />

Grouped positive regions<br />

Mean (median) N Mean (median) N<br />

(LFC LCL) depth 1560 (1590) m 24 740 (720) m 24<br />

0–3 km AGL shear 10.7 (9.0) m s 1 24 14.7 (14.9) m s 1 24<br />

Freez<strong>in</strong>g level (FL) 4070 (4100) m AGL 24 3780 (3770) m AGL 24<br />

CAPE between LFC and 10°C level 397 (385) J kg 1 24 199 (202) J kg 1 19<br />

Surface temperature (T) 26.5 (27.4)°C 24 31.0 (29.6)°C 24<br />

In agreement with Knapp (1994), negative s<strong>to</strong>rms occurred<br />

<strong>in</strong> a moister environment than positive s<strong>to</strong>rms as<br />

<strong>in</strong>dicated by significantly higher mean low-level mix<strong>in</strong>g<br />

ratio, PW, and surface dewpo<strong>in</strong>t, along with a significantly<br />

lower mean surface dewpo<strong>in</strong>t depression. The<br />

difference <strong>in</strong> means for all <strong>of</strong> the above moisture parameters<br />

was very highly significant (99.9% level) as<br />

shown <strong>in</strong> Table 4. Negative s<strong>to</strong>rms also occurred <strong>in</strong><br />

regions <strong>of</strong> noticeably higher midlevel relative humidity,<br />

although the differences <strong>in</strong> the means were significant<br />

only at the 90% level (Table 7).<br />

The positive mesoscale regions were characterized by<br />

a significantly higher mean LCL as suggested by W05.<br />

In fact, the mean LCL for positive regions (2080 m<br />

AGL) was 1.9 times greater than that for negative regions<br />

(1120 m AGL). A higher LCL <strong>in</strong> comb<strong>in</strong>ation<br />

with a slightly lower mean freez<strong>in</strong>g level (and wet-bulb<br />

zero height) resulted <strong>in</strong> a much shallower mean WCD<br />

<strong>in</strong> positive mesoscale regions as also postulated by<br />

W05. More specifically, the WCD was 1.3 km deeper <strong>in</strong><br />

negative s<strong>to</strong>rms (2950 m) as compared <strong>to</strong> positive<br />

s<strong>to</strong>rms (1700 m). The difference <strong>in</strong> mean LCL and<br />

WCD between the two regions was very highly significant<br />

(99.9% level; Table 4). By comparison, the LFC <strong>in</strong><br />

positive and negative s<strong>to</strong>rm environments was not statistically<br />

different (mean and median <strong>of</strong> about 3 km for<br />

both; Table 8).<br />

Mean lapse rates <strong>in</strong> the low <strong>to</strong> midtroposphere (850–<br />

500 and 700–500 hPa) were steeper <strong>in</strong> positive regions<br />

(Tables 4 and 6). The mean surface temperature was<br />

greater <strong>in</strong> positive regions as well (Table 5). The mean<br />

EL was higher <strong>in</strong> negative regions, and thus, despite<br />

little difference <strong>in</strong> the LFC between negative and positive<br />

regions, the depth <strong>of</strong> the free convective layer<br />

(EL LFC) was greater <strong>in</strong> negative regions (Table 6).<br />

Interest<strong>in</strong>gly, there was no significant difference <strong>in</strong><br />

the mean <strong>to</strong>tal CAPE (LFC–EL) or mean lifted <strong>in</strong>dex<br />

(LI) for positive and negative s<strong>to</strong>rms (Table 8). The<br />

mean CAPE was moderate (roughly 2000 J kg 1 ) and<br />

the mean lifted <strong>in</strong>dex was low (6° <strong>to</strong> 7°C), suggest<strong>in</strong>g<br />

very unstable air masses and ample <strong>in</strong>stability for<br />

strong updraft development <strong>in</strong> both regions. There was<br />

also no significant difference <strong>in</strong> the mean deep-layer<br />

(0–6 km AGL) vertical w<strong>in</strong>d shear for positive and<br />

negative s<strong>to</strong>rms (Table 8), as also found by Reap and<br />

MacGorman (1989) and Curran and Rust (1992).<br />

Therefore, it is not surpris<strong>in</strong>g that the BRN did not<br />

differ significantly between positive and negative mesoscale<br />

regions either (Table 8). However, the mean<br />

convective <strong>in</strong>hibition (CIN) was significantly greater<br />

for negative regions (Table 6). Given recent <strong>in</strong>terest <strong>in</strong><br />

the relationship between dom<strong>in</strong>ant CG lightn<strong>in</strong>g polarity<br />

and position <strong>of</strong> the e ridge (e.g., Smith et al. 2000;<br />

CRP03), it is <strong>in</strong>terest<strong>in</strong>g (but perhaps not surpris<strong>in</strong>g as<br />

discussed <strong>in</strong> section 5b) <strong>to</strong> note that mean e values did<br />

not differ significantly between negative and positive<br />

regions, and were <strong>in</strong> fact very similar (Table 8).<br />

Mean CAPE and NCAPE were calculated for sev-<br />

TABLE 6. Same as <strong>in</strong> Table 4, but the difference <strong>in</strong> each mean (median) sound<strong>in</strong>g parameter was significant<br />

(95% level, 0.01 p 0.05).<br />

Significant: 95% level<br />

Grouped negative regions<br />

Grouped positive regions<br />

Mean (median) N Mean (median) N<br />

CIN 67 (43) J kg 1 24 26 (10) J kg 1 24<br />

EL 12500 (12600) m AGL 22 11700 (11900) m AGL 10<br />

700–500-hPa lapse rate 7.7 (7.9)°C km 1 24 8.4 (8.3)°C km 1 18<br />

0–2 km AGL s<strong>to</strong>rm-relative w<strong>in</strong>d speed 7.0 (6.2) m s 1 22 10.2 (9.8) m s 1 13<br />

Depth <strong>of</strong> free convective layer (EL LFC) 9810 (10 090) m 22 8600 (8710) m 10<br />

CAPE between 10° and 40°C levels 957 (1010) J kg 1 22 1210 (1270) J kg 1 10<br />

NCAPE between LFC and 40°C level 0.19 (0.20) m s 2 22 0.24 (0.25) m s 2 10


APRIL 2007 C A R E Y A N D B U F F A L O 1337<br />

TABLE 7. Same as <strong>in</strong> Table 4, but the difference <strong>in</strong> each mean (median) sound<strong>in</strong>g parameter was <strong>in</strong>significant at the 95% level but<br />

significant at the 90% level (0.05 p 0.1).<br />

Significance level: 90%<br />

Grouped negative regions<br />

Grouped positive regions<br />

Mean (median) N Mean (median) N<br />

0–3 km AGL SREH 72 (67) m 2 s 2 22 163 (119) m 2 s 2 13<br />

Midlevel relative humidity (700–500-hPa layer) 42 (45) % 24 32 (28) % 18<br />

CAPE between LFC and 0°C level 137 (152) J kg 1 24 57 (43) J kg 1 24<br />

NCAPE (LFC–EL) 0.19 (0.19) m s 2 22 0.22 (0.22) m s 2 10<br />

NCAPE between 10° and 40°C levels 0.24 (0.25) m s 2 22 0.29 (0.31) m s 2 10<br />

eral different vertical layers. In general, there was more<br />

CAPE at warmer temperatures (LFC <strong>to</strong> 10°C) <strong>in</strong><br />

negative regions and more CAPE at colder temperatures<br />

(10° <strong>to</strong> 40°C) <strong>in</strong> positive regions (Tables 5–7).<br />

Inconsistent with the results from GW02, mean CAPE<br />

between the LFC and 0°C level and between the LFC<br />

and the 10°C level were significantly greater for negative<br />

s<strong>to</strong>rms (Tables 7 and 5, respectively). However,<br />

mean NCAPE between the LFC and 10°C level was<br />

not significantly different between negative and positive<br />

regions (Table 8), <strong>in</strong>dicat<strong>in</strong>g that the greater CAPE<br />

values for negative s<strong>to</strong>rms <strong>in</strong> these lower layers was due<br />

<strong>to</strong> deeper mean LFC <strong>to</strong> 0°C and LFC <strong>to</strong> 10°C layers<br />

<strong>in</strong> negative regions, rather than <strong>to</strong> greater mean parcel<br />

buoyant acceleration with<strong>in</strong> these layers. Mean CAPE<br />

between the 10° and 40°C levels and mean NCAPE<br />

between the LFC and 40°C levels were both 26%<br />

greater <strong>in</strong> positive regions, with the difference <strong>in</strong> means<br />

for both parameters significant at the 95% level (Table<br />

6). Mean NCAPE between the 10° and 40°C levels<br />

was also greater for positive s<strong>to</strong>rms but was only significant<br />

at the 90% level (Table 7). Mean <strong>to</strong>tal NCAPE<br />

(LFC–EL) was greater for positive s<strong>to</strong>rms but the difference<br />

<strong>in</strong> means was only significant at the 90% level<br />

(Table 7). Similar <strong>to</strong> <strong>to</strong>tal CAPE, CAPE between the<br />

LFC and the height <strong>of</strong> the 40°C level was not significantly<br />

different between negative and positive s<strong>to</strong>rms<br />

(Table 8). In summary, CAPE and more precisely mean<br />

buoyancy acceleration (NCAPE) was larger for positive<br />

s<strong>to</strong>rms with<strong>in</strong> the mixed-phase zone (0° <strong>to</strong> 40°C),<br />

which is critical for NIC and lightn<strong>in</strong>g production.<br />

Consistent with the results <strong>of</strong> GW02, the 0–3 km<br />

AGL vertical w<strong>in</strong>d shear was significantly stronger <strong>in</strong><br />

positive (14.7 m s 1 ) versus negative (10.7 m s 1 ) mesoscale<br />

regions and hence could imply stronger dynamical<br />

updraft forc<strong>in</strong>g <strong>of</strong> positive s<strong>to</strong>rms. The difference <strong>in</strong><br />

the mean 0–3-km shear for the positive and negative<br />

regions was highly significant (99% level; Table 5).<br />

However, as noted above, 0–6 km AGL shear did not<br />

differ significantly between negative and positive<br />

s<strong>to</strong>rms, nor did 0–2 km AGL shear (Table 8). Low-level<br />

(0–2 km AGL) s<strong>to</strong>rm-relative w<strong>in</strong>d speed was significantly<br />

higher <strong>in</strong> positive regions (Table 6). Low-level<br />

s<strong>to</strong>rm-relative w<strong>in</strong>d speeds are a proxy for low-level<br />

TABLE 8. Same as <strong>in</strong> Table 4, but the difference <strong>in</strong> each mean (median) sound<strong>in</strong>g parameter was <strong>in</strong>significant at the 90% level<br />

(p 0.1).<br />

Not significant at the 90% level<br />

Grouped negative regions<br />

Grouped positive regions<br />

Mean (median) N Mean (median) N<br />

CAPE (LFC–EL) 1920 (2030) J kg 1 22 1950 (2020) J kg 1 10<br />

LI 6.9 (6.0)°C 24 6.1 (6.5)°C 18<br />

LFC 2680 (2790) m AGL 24 2820 (3070) m AGL 24<br />

Mean relative humidity (through full depth <strong>of</strong> sound<strong>in</strong>g) 37 (33) % 22 29 (26) % 10<br />

4–6 km AGL s<strong>to</strong>rm-relative w<strong>in</strong>d speed 10.7 (9.8) m s 1 22 10.6 (9.3) m s 1 13<br />

6–10 km AGL s<strong>to</strong>rm-relative w<strong>in</strong>d speed 15.8 (14.9) m s 1 22 15.1 (14.4) m s 1 10<br />

9–11 km AGL s<strong>to</strong>rm-relative w<strong>in</strong>d speed 22.8 (20.8) m s 1 22 19.4 (18.8) m s 1 10<br />

S<strong>to</strong>rm-relative w<strong>in</strong>d speed at EL 25.8 (24.9) m s 1 22 21.3 (22.4) m s 1 10<br />

0–2 km AGL shear 8.2 (7.7) m s 1 24 9.2 (8.5) m s 1 24<br />

0–6 km AGL shear 17.7 (17.5) m s 1 22 18.5 (18.5) m s 1 13<br />

BRN 149 (82) 22 91 (49) 10<br />

EHI (us<strong>in</strong>g 0–3 km AGL SREH) 0.8 (0.7) 22 2.0 (1.4) 10<br />

NCAPE between LFC and 10°C level 0.13 (0.12) m s 2 24 0.11 (0.12) m s 2 19<br />

CAPE between LFC and 40°C level 1340 (1340) J kg 1 22 1410 (1440) J kg 1 10<br />

Equivalent potential temperature ( e ) 73.2 (73.0)°C 24 71.9 (74.1)°C 24


1338 MONTHLY WEATHER REVIEW VOLUME 135<br />

FIG. 2. Relative frequency his<strong>to</strong>gram <strong>of</strong> LCL (m, AGL) for<br />

<strong>in</strong>dividual sound<strong>in</strong>gs associated with negative and positive s<strong>to</strong>rms.<br />

Labels along the horizontal axis represent the maximum value for<br />

the b<strong>in</strong>.<br />

<strong>in</strong>flow strength. Model<strong>in</strong>g studies have shown that the<br />

low-level outflow strength is detrimental <strong>to</strong> supercell<br />

ma<strong>in</strong>tenance and <strong>in</strong>tensity when it is <strong>to</strong>o strong relative<br />

<strong>to</strong> the low-level <strong>in</strong>flow (e.g., Weisman and Klemp 1982;<br />

Brooks et al. 1994a) because it undercuts the warm<br />

<strong>in</strong>flow <strong>in</strong><strong>to</strong> the updraft thereby weaken<strong>in</strong>g the convection.<br />

As a result, stronger <strong>in</strong>flow may allow the sustenance<br />

<strong>of</strong> positive s<strong>to</strong>rms <strong>in</strong> the presence <strong>of</strong> strong outflow.<br />

Interest<strong>in</strong>gly, s<strong>to</strong>rm-relative w<strong>in</strong>d speeds <strong>in</strong> the<br />

mid- and upper troposphere (4–6, 6–10, and 9–11 km<br />

AGL; EL) were fairly similar between negative and<br />

positive regions (Table 8), despite play<strong>in</strong>g a hypothesized<br />

role <strong>in</strong> supercell microphysics and dynamics<br />

(e.g., Brooks et al. 1994a; Rasmussen and Straka 1998).<br />

The mean 0–3 km AGL s<strong>to</strong>rm relative helicity (SREH)<br />

was over twice as high <strong>in</strong> positive regions than <strong>in</strong> negative<br />

regions. However, due <strong>to</strong> a large variation <strong>in</strong> 0–3-<br />

km SREH values, this difference <strong>in</strong> means was only<br />

significant at the 90% level (Table 7). SREH provides<br />

a means <strong>of</strong> assess<strong>in</strong>g the tendency for mesocyclone formation<br />

<strong>in</strong> supercells (Davies-Jones et al. 1990) and<br />

hence dynamical forc<strong>in</strong>g <strong>of</strong> the updraft (e.g., Weisman<br />

and Klemp 1982). The 0–3-km shear and SREH results<br />

suggest that positive s<strong>to</strong>rms apparently experienced<br />

stronger dynamical forc<strong>in</strong>g <strong>of</strong> the updraft than negative<br />

s<strong>to</strong>rms. S<strong>in</strong>ce <strong>to</strong>tal CAPE was so similar for positive<br />

and negative regions, the energy helicity <strong>in</strong>dex (EHI;<br />

us<strong>in</strong>g 0–3 km AGL SREH) did not differ significantly<br />

between positive and negative regions (Table 8).<br />

b. Variability <strong>of</strong> sound<strong>in</strong>g parameters<br />

Relative frequency his<strong>to</strong>grams <strong>of</strong> select sound<strong>in</strong>g parameters<br />

that characterize the mesoscale environments<br />

FIG. 3. Same as <strong>in</strong> Fig. 2, but for WCD (m).<br />

<strong>of</strong> negative and positive s<strong>to</strong>rms from <strong>in</strong>dividual sound<strong>in</strong>gs<br />

(shown <strong>in</strong> Figs. 2–11) highlight the significant (i.e.,<br />

significance level 95% or p 0.05) differences <strong>in</strong><br />

these variables and yet assess the degree <strong>of</strong> overlap <strong>in</strong><br />

environmental conditions between the two types <strong>of</strong> regions.<br />

F<strong>in</strong>ally, the his<strong>to</strong>grams demonstrate the range <strong>of</strong><br />

variability <strong>in</strong> each parameter associated with both negative<br />

and positive s<strong>to</strong>rms.<br />

In agreement with the mean value comparisons and<br />

significance tests the his<strong>to</strong>grams (shown <strong>in</strong> Figs. 2–11)<br />

make it readily apparent that LCL (Fig. 2) and especially<br />

WCD (Fig. 3) differed much more between positive<br />

and negative mesoscale regions than most other<br />

parameters <strong>in</strong>vestigated. Although the populations<br />

were not completely dist<strong>in</strong>ct, there was relatively little<br />

overlap, especially for WCD. The modal LCL for positive<br />

and negative s<strong>to</strong>rms was 2000 and 1000 m AGL,<br />

respectively. The WCD for positive s<strong>to</strong>rms was somewhat<br />

bimodal with relative maxima around 1300 m,<br />

which was the primary peak, and 2100 m while the<br />

mode for negative s<strong>to</strong>rms was 3100 m. There was one<br />

outlier sound<strong>in</strong>g launched <strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> negative<br />

s<strong>to</strong>rms that was characterized by an LCL <strong>of</strong> approximately<br />

2500 m AGL and a WCD <strong>of</strong> about 1100 m.<br />

While high LCL and shallow WCD were not exclusively<br />

associated with positive s<strong>to</strong>rms, the degree <strong>of</strong><br />

separation between positive and negative regions evident<br />

<strong>in</strong> the LCL and WCD <strong>of</strong> <strong>in</strong>dividual sound<strong>in</strong>gs was<br />

noteworthy. As such, these results strongly support a<br />

role for the LCL and WCD <strong>in</strong> <strong>in</strong>fluenc<strong>in</strong>g the CG lightn<strong>in</strong>g<br />

polarity <strong>of</strong> severe s<strong>to</strong>rms <strong>in</strong> the central CONUS as<br />

first suggested by W05.<br />

Figures 4–6 illustrate once aga<strong>in</strong> that negative s<strong>to</strong>rms<br />

occurred <strong>in</strong> more moist environments than positive<br />

s<strong>to</strong>rms similar <strong>to</strong> the suggestion by Knapp (1994). The<br />

moister negative s<strong>to</strong>rm environments were noticeable


APRIL 2007 C A R E Y A N D B U F F A L O 1339<br />

FIG. 4. Same as <strong>in</strong> Fig. 2, but for surface dewpo<strong>in</strong>t depression<br />

(°C).<br />

FIG. 6. Same as <strong>in</strong> Fig. 2, but for PW (cm) between the surface<br />

and 400 hPa.<br />

<strong>in</strong> measurements <strong>of</strong> surface moisture (Fig. 4), low-level<br />

moisture (Fig. 5), and low- <strong>to</strong> midtropospheric moisture<br />

(Fig. 6). The moisture parameters exhibited more overlap<br />

<strong>of</strong> <strong>in</strong>dividual sound<strong>in</strong>g values between negative and<br />

positive s<strong>to</strong>rms than did LCL and WCD, but the respective<br />

modes for positive and negative s<strong>to</strong>rms <strong>of</strong> all<br />

three moisture parameters were clearly dist<strong>in</strong>ct. The<br />

same can be said for 0–3 km AGL shear (Fig. 7). The<br />

modes were obviously different (7 m s 1 for negative<br />

s<strong>to</strong>rms versus 15 m s 1 for positive s<strong>to</strong>rms), with the<br />

positive sound<strong>in</strong>g population occupy<strong>in</strong>g the higher end<br />

<strong>of</strong> the parameter spectrum, and the negative sound<strong>in</strong>g<br />

population occupy<strong>in</strong>g the lower end, confirm<strong>in</strong>g the<br />

f<strong>in</strong>d<strong>in</strong>gs <strong>of</strong> GW02.<br />

Look<strong>in</strong>g at 0–2 km AGL s<strong>to</strong>rm-relative w<strong>in</strong>d speed<br />

(Fig. 8), the separation between sound<strong>in</strong>g populations<br />

<strong>of</strong> negative and positive regions was less def<strong>in</strong>ed, as is<br />

expected based on the decreas<strong>in</strong>g significance level <strong>of</strong><br />

the difference <strong>in</strong> means from Tables 4 <strong>to</strong> 6. The negative<br />

sound<strong>in</strong>g population was bimodal, with one mode (9.5<br />

ms 1 ) greater than and the other mode (3.5 m s 1 ) less<br />

than the positive sound<strong>in</strong>g mode (6.5 m s 1 ). The positive<br />

sound<strong>in</strong>g population was skewed <strong>to</strong>ward higher<br />

s<strong>to</strong>rm-relative w<strong>in</strong>d speed values up <strong>to</strong> 17 m s 1 . While<br />

there was more overlap for this parameter, the samples<br />

for the positive and negative regions were still dist<strong>in</strong>ct.<br />

F<strong>in</strong>ally, parameters related <strong>to</strong> buoyant or conditional<br />

<strong>in</strong>stability were considered. The 850–500-hPa lapse rate<br />

sample for each region was dist<strong>in</strong>ct with negative (positive)<br />

region values skewed <strong>to</strong>ward lower (higher) values<br />

(Fig. 9). Although there was some overlap <strong>in</strong> the<br />

lapse rate samples between 6.9° and 8.4°C km 1 ,an<br />

overwhelm<strong>in</strong>g majority <strong>of</strong> negative (positive) region<br />

lapse rates were less (greater) than 7.2°C km 1 . Rela-<br />

FIG. 5. Same as <strong>in</strong> Fig. 2, but for mean mix<strong>in</strong>g ratio (g kg 1 )<strong>in</strong><br />

the lowest 100 hPa.<br />

FIG. 7. Same as <strong>in</strong> Fig. 2, but for low-level (0–3 km AGL)<br />

vertical w<strong>in</strong>d shear (m s 1 ).


1340 MONTHLY WEATHER REVIEW VOLUME 135<br />

FIG. 8. Same as <strong>in</strong> Fig. 2, but for 0–2 km AGL s<strong>to</strong>rm-relative<br />

w<strong>in</strong>d speed (m s 1 ).<br />

FIG. 10. Same as <strong>in</strong> Fig. 2, but for CAPE (J kg 1 ) between the<br />

10° and 40°C levels.<br />

tive frequency his<strong>to</strong>grams <strong>of</strong> CAPE between 10° and<br />

40°C and NCAPE between the LFC and 40°C are<br />

presented <strong>in</strong> Figs. 10 and 11, respectively. There was<br />

noticeable overlap between negative and positive region<br />

CAPE values calculated between the 10° <strong>to</strong><br />

40°C levels with close but dist<strong>in</strong>ct modes (negative:<br />

1100 J kg 1 ; positive: 1300 J kg 1 ). There was less overlap<br />

<strong>in</strong> the samples <strong>of</strong> NCAPE from the LFC <strong>to</strong> the<br />

40°C level for the two regions (negative mode: 0.21<br />

ms 2 ; positive mode: 0.26 m s 2 ). CAPE and NCAPE<br />

frequency his<strong>to</strong>grams calculated for other layers <strong>of</strong> the<br />

atmosphere considered <strong>in</strong> this study (not shown but see<br />

Tables 7 and 8 for means) were characterized by significantly<br />

more overlap than Figs. 10 and 11. Indeed,<br />

the samples <strong>of</strong> traditional (i.e., LFC–EL) CAPE calculated<br />

from sound<strong>in</strong>gs <strong>in</strong> the negative and positive mesoscale<br />

regions were <strong>in</strong>dist<strong>in</strong>guishable. In summary,<br />

positive s<strong>to</strong>rms generally formed <strong>in</strong> regions characterized<br />

by larger 850–500-hPa lapse rates and larger<br />

NCAPE (i.e., mean parcel acceleration associated with<br />

buoyancy; Blanchard 1998) up <strong>to</strong> the <strong>to</strong>p <strong>of</strong> the mixedphase<br />

zone (i.e., 40°C).<br />

c. Median environmental conditions associated with<br />

the mesoscale regions<br />

To carefully <strong>in</strong>vestigate the hypothesis that regional<br />

CG lightn<strong>in</strong>g polarity is controlled by systematic differences<br />

<strong>in</strong> mesoscale environmental conditions (e.g.,<br />

MacGorman and Burgess 1994), median values <strong>of</strong> select<br />

parameters, which were found <strong>to</strong> differ significantly<br />

between negative and positive regions <strong>in</strong> the overall<br />

comparisons (Tables 4–6), were calculated for each <strong>of</strong><br />

the n<strong>in</strong>e mesoscale regions separately (Table 9 and<br />

Figs. 12 and 13). As was the case with the overall com-<br />

FIG. 9. Same as <strong>in</strong> Fig. 2, but for 850–500-hPa lapse rate<br />

(°C km 1 ).<br />

FIG. 11. Same as <strong>in</strong> Fig. 2, but for NCAPE (m s 2 ) between the<br />

LFC and 40°C level.


APRIL 2007 C A R E Y A N D B U F F A L O 1341<br />

TABLE 9. Median parameter values for the n<strong>in</strong>e <strong>in</strong>dividual mesoscale regions <strong>in</strong>vestigated. Parameters listed were found <strong>to</strong> differ<br />

significantly (i.e., significance level 95% or p 0.05) between negative and positive regions <strong>in</strong> the overall group comparisons (Tables<br />

4–6). However, only median values are directly compared here due <strong>to</strong> the small sample size for each <strong>in</strong>dividual region.<br />

Negative mesoscale regions<br />

Positive mesoscale regions<br />

23 May 24 May 15 Jun 4 Jun 12 Jun 23 May 24 May 15 Jun 19 Jun<br />

WCD (m) 3140 2930 2600 3090 3110 1230 1780 1970 1160<br />

LCL (m AGL) 810 800 1570 1040 1190 1670 1720 1950 2920<br />

Mean mix<strong>in</strong>g ratio <strong>in</strong> the lowest 12.1 12.4 10.9 14.1 16.3 9.9 10.7 11.4 10.0<br />

100 hPa (g kg 1 )<br />

PW, surface <strong>to</strong> 400 hPa (cm) 3.5 3.4 3.4 3.5 4.2 2.4 2.1 2.9 2.6<br />

Surface dewpo<strong>in</strong>t depression (°C) 5.1 4.4 10.7 7.3 8.8 11.0 12.6 15.0 23.5<br />

850–500-hPa lapse rate (°C km 1 ) 6.5 6.8 6.9 7.1 7.3 8.5 7.9 8.2 9.1<br />

0–3 km AGL shear (m s 1 ) 9.3 7.7 17.0 6.7 13.1 16.0 15.7 16.5 11.6<br />

0–2 km AGL s<strong>to</strong>rm-relative<br />

10.0 18.0 22.0 7.0 19.5 20.0 13.0 26.0 16.0<br />

w<strong>in</strong>d speed (m s 1 )<br />

NCAPE, LFC <strong>to</strong> 40°C (ms 2 ) 0.12 0.18 0.10 0.19 0.23 0.22 0.16 0.25 0.26<br />

CAPE, 10° <strong>to</strong> 40°C (Jkg 1 ) 546 894 460 1010 1090 1100 770 1320 1300<br />

parisons, WCD and LCL were the strongest discrim<strong>in</strong>a<strong>to</strong>rs<br />

between negative and positive mesoscale regions.<br />

The LCL values for all <strong>in</strong>dividual positive and<br />

negative regions were dist<strong>in</strong>ct with values greater than<br />

and less than 1600 m AGL, respectively (Table 9 and<br />

Fig. 12). The median WCD values for <strong>in</strong>dividual positive<br />

regions were all less than 2000 m, while the correspond<strong>in</strong>g<br />

median values for negative regions were all<br />

greater than 2500 m (Table 9 and Fig. 13). Negative<br />

mesoscale regions were once aga<strong>in</strong> noticeably moister<br />

than positive mesoscale regions, as <strong>in</strong>dicated by the median<br />

mix<strong>in</strong>g ratio <strong>in</strong> the lowest 100 hPa, the PW between<br />

the surface and 400 hPa, and the surface dewpo<strong>in</strong>t<br />

depression (Table 9). Median PW <strong>in</strong> negative regions<br />

was 3.3 cm while it was 3.0 cm <strong>in</strong> positive<br />

regions. Median surface dewpo<strong>in</strong>t depression was consistently<br />

11°C <strong>in</strong> negative regions and 11°C <strong>in</strong> positive<br />

regions, consistent with lower and higher LCLs <strong>in</strong><br />

each region, respectively (W05). There was slight overlap<br />

<strong>in</strong> the median low-level mix<strong>in</strong>g ratios between the<br />

two types <strong>of</strong> regions. Lapse rates between 850 and 500<br />

hPa proved <strong>to</strong> be consistently higher <strong>in</strong> positive mesoscale<br />

regions. Median lapse rates <strong>in</strong> the 850–500-hPa<br />

layer were greater (less) than 7.5°C km 1 for all <strong>in</strong>dividual<br />

positive (negative) mesoscale regions (Table 9<br />

and Fig. 13). Although grouped positive mesoscale regions<br />

were generally characterized by larger 0–3 km<br />

AGL shear, 0–2 km AGL s<strong>to</strong>rm-relative w<strong>in</strong>d speed,<br />

NCAPE (LFC <strong>to</strong> 40°C), and CAPE (10° <strong>to</strong> 40°C)<br />

when compared <strong>to</strong> grouped negative mesoscale regions<br />

(Tables 5 and 6), there was considerable overlap be-<br />

FIG. 12. Scatterplot <strong>of</strong> the median NCAPE between the LFC<br />

and 40°C level vs the height <strong>of</strong> the LCL for each <strong>of</strong> the n<strong>in</strong>e<br />

mesoscale regions. The size <strong>of</strong> each bubble <strong>in</strong> the scatterplot is<br />

proportional <strong>to</strong> the median 0–2 km AGL s<strong>to</strong>rm-relative w<strong>in</strong>d<br />

speed <strong>in</strong> each region, which is also <strong>in</strong>dicated by the label on each<br />

bubble. The 23 May negative mesoscale region, which was the<br />

only region not characterized by widespread severe s<strong>to</strong>rms, is<br />

labeled as “nonsevere.”<br />

FIG. 13. Same as <strong>in</strong> Fig. 12, but for median 850–500-hPa lapse<br />

rate vs median WCD for each <strong>in</strong>dividual mesoscale region. The<br />

bubble size (and labels) represent the median 0–3 km AGL shear<br />

for each region.


1342 MONTHLY WEATHER REVIEW VOLUME 135<br />

tween the median parameter values <strong>of</strong> <strong>in</strong>dividual negative<br />

and positive mesoscale regions (Table 9 and Figs.<br />

12 and 13).<br />

The contribution <strong>of</strong> each parameter <strong>to</strong> modulat<strong>in</strong>g<br />

updraft strength and supercooled liquid water content<br />

<strong>in</strong> each region may not be equal, possibly expla<strong>in</strong><strong>in</strong>g<br />

some <strong>of</strong> the overlap. One parameter may compensate<br />

for another. For example, on 15 June the positive region<br />

LCL (WCD) was only slightly higher (smaller)<br />

than the <strong>in</strong> the negative region on the same day. However,<br />

the positive region CAPE between 10° and<br />

40°C (NCAPE from LFC <strong>to</strong> 40°C) was 2.9 (2.5)<br />

times larger than <strong>in</strong> the negative region on 15 June such<br />

that the updraft was apparently still stronger. F<strong>in</strong>ally, it<br />

is important <strong>to</strong> recall that the sample size for each <strong>in</strong>dividual<br />

region is relatively small so that regional differences<br />

(or lack there<strong>of</strong>) must be viewed with some<br />

caution.<br />

A key f<strong>in</strong>d<strong>in</strong>g <strong>of</strong> earlier case studies was the observation<br />

that s<strong>to</strong>rms on the same day pass<strong>in</strong>g over similar<br />

mesoscale regions produced similar CG lightn<strong>in</strong>g behavior<br />

(e.g., Branick and Doswell 1992; MacGorman<br />

and Burgess 1994; Smith et al. 2000). To <strong>in</strong>vestigate this<br />

issue, we compared the environmental conditions <strong>in</strong><br />

positive and negative s<strong>to</strong>rm regions occurr<strong>in</strong>g adjacent<br />

<strong>to</strong> each other on the same day (e.g., we compared the<br />

parameters <strong>of</strong> the 23 May positive region <strong>to</strong> those <strong>of</strong> the<br />

23 May negative region shown <strong>in</strong> Table 9). Opportunities<br />

for a daily comparison <strong>of</strong> positive and negative mesoscale<br />

regions were available on 23 and 24 May and 15<br />

June. The median positive and negative region values<br />

<strong>of</strong> WCD, LCL, PW, dewpo<strong>in</strong>t depression, and 850–500-<br />

hPa lapse rate were significantly different on a daily<br />

basis (Table 9). The daily regional differences for these<br />

parameters were also consistent with the overall<br />

grouped results (e.g., Tables 4–6). When consider<strong>in</strong>g<br />

the rest <strong>of</strong> the environmental parameters <strong>in</strong> Table 9, the<br />

daily comparison <strong>of</strong> positive and negative mesoscale<br />

regions was somewhat mixed and not always consistent<br />

with the overall grouped results. For example, the median<br />

low-level mix<strong>in</strong>g ratio was somewhat larger and<br />

the 0–3-km shear was somewhat smaller <strong>in</strong> the positive<br />

region on 15 June. Similarly, the median 0–2-km s<strong>to</strong>rmrelative<br />

w<strong>in</strong>d speed, NCAPE (LFC <strong>to</strong> 40°C), and<br />

CAPE (10° <strong>to</strong> 40°C) were all larger <strong>in</strong> the negative<br />

mesoscale region on 24 May. On the other hand, all<br />

<strong>of</strong> the medians listed <strong>in</strong> Table 9 for 23 May were dist<strong>in</strong>ctly<br />

different between the positive and negative mesoscale<br />

regions and consistent with the overall grouped<br />

results.<br />

Sometimes the daily differences <strong>in</strong> medians between<br />

the positive and negative regions were quite dramatic<br />

(Table 9). On 23 and 24 May, the median WCD was 2.6<br />

and 1.7 times smaller, the median LCL was 2.1 and 2.2<br />

times higher, the median surface dewpo<strong>in</strong>t depression<br />

was 2.2 and 2.9 times larger, and the median 0–3-km<br />

shear was 1.7 and 2.0 times larger <strong>in</strong> the positive mesoscale<br />

region, respectively. The median 0–2-km s<strong>to</strong>rmrelative<br />

w<strong>in</strong>d speed was 2.0 times larger <strong>in</strong> the positive<br />

region on 23 May. On 15 June, the median NCAPE<br />

(LFC <strong>to</strong> 40°C) and CAPE (10° <strong>to</strong> 40°C) values<br />

were 2.5 and 2.9 times larger <strong>in</strong> the positive region,<br />

respectively.<br />

d. Intraregional variation <strong>of</strong> environmental<br />

condition—The positive region <strong>of</strong> 24 May 2002<br />

Dropsondes released from the FI Learjet on 24 May<br />

permitted the <strong>in</strong>traregional <strong>in</strong>vestigation <strong>of</strong> a positive<br />

s<strong>to</strong>rm environment at a very high spatial and temporal<br />

resolution. The dropsondes were released <strong>in</strong> the eastern<br />

Texas panhandle and southwest Oklahoma (Fig.<br />

14). The dropsonde l<strong>in</strong>e was approximately perpendicular<br />

<strong>to</strong> the surface boundaries associated with convective<br />

<strong>in</strong>itiation (CI) and near-surface e ridge, and<br />

roughly parallel <strong>to</strong> subsequent s<strong>to</strong>rm motion (not<br />

shown). 4 The numbers <strong>of</strong> positive and negative CG<br />

lightn<strong>in</strong>g flashes dur<strong>in</strong>g the 6-h period were accumulated<br />

<strong>in</strong> 25 km 2 grid boxes over the region <strong>of</strong> <strong>in</strong>terest.<br />

From this gridded data, we computed the fraction <strong>of</strong><br />

positive CG flashes over all CG flashes. S<strong>in</strong>ce the<br />

s<strong>to</strong>rms <strong>in</strong>itiated just <strong>to</strong> the east <strong>of</strong> the dry l<strong>in</strong>e and generally<br />

moved eastward, Fig. 14 essentially provides a<br />

view <strong>of</strong> the evolution <strong>of</strong> the positive CG fraction dur<strong>in</strong>g<br />

the life cycle <strong>of</strong> the event. As shown <strong>in</strong> Fig. 14, CG<br />

lightn<strong>in</strong>g polarity transitioned from positive <strong>to</strong> negative<br />

as s<strong>to</strong>rms <strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> the 24 May dropsondes<br />

moved from west <strong>to</strong> east across Oklahoma and away<br />

from the dryl<strong>in</strong>e. Although the dropsondes were<br />

dropped entirely <strong>in</strong> the positive mesoscale region (Fig.<br />

14), the CG percentage generally decreased from<br />

west <strong>to</strong> east with<strong>in</strong> the identified positive region (Figs.<br />

14 and 15).<br />

Parameters that were found <strong>to</strong> differ significantly and<br />

consistently between negative and positive regions <strong>in</strong><br />

the grouped and regional comparisons above were calculated<br />

for all <strong>of</strong> the good (i.e., conta<strong>in</strong><strong>in</strong>g no bad data)<br />

24 May dropsondes from 2022 <strong>to</strong> 2046 UTC as shown <strong>in</strong><br />

Fig. 15a (NCAPE and 0–3 km AGL shear) and Fig. 15b<br />

4 The <strong>in</strong>itiation <strong>of</strong> deep convection <strong>in</strong> Figs. 14 and 15 was determ<strong>in</strong>ed<br />

by manual <strong>in</strong>spection <strong>of</strong> the Geostationary Operational<br />

<strong>Environmental</strong> Satellite (GOES-8) <strong>in</strong>frared (IR) and visible imagery.<br />

The first CG lightn<strong>in</strong>g (<strong>in</strong>clud<strong>in</strong>g positive flashes) was with<strong>in</strong><br />

roughly 15–30 km eastward <strong>of</strong> the dryl<strong>in</strong>e and CI.


APRIL 2007 C A R E Y A N D B U F F A L O 1343<br />

FIG. 14. Percent positive flashes (%, shaded as shown) for 2000 UTC 24 May 2002–0200 UTC 25 May 2002. Black diamonds <strong>in</strong>dicate<br />

dropsonde locations (from left <strong>to</strong> right, dropsonde release times were 2022, 2025, 2031, 2034, 2037, 2043, and 2046 UTC). Dropsondes<br />

released at 2028 and 2040 UTC conta<strong>in</strong>ed bad data and thus were not used. Orange rectangles (north–south dimension 80 km,<br />

east–west dimension 27.4 km) centered on dropsonde locations denote the areas for which CG flash characteristics (number <strong>of</strong><br />

negative flashes, number <strong>of</strong> positive flashes, and percent <strong>of</strong> positive flashes) were determ<strong>in</strong>ed. The area <strong>to</strong> the north and west <strong>of</strong> the<br />

dotted l<strong>in</strong>e is the positive mesoscale region. The plot is centered over the eastern Texas panhandle, north-central Texas, and southwest<br />

Oklahoma.<br />

5 Dropsondes were sometimes dropped from pressures higher<br />

than 500 hPa. To ma<strong>in</strong>ta<strong>in</strong> consistency for all lapse rates calculated<br />

from dropsondes, we slightly modified the lower pressure <strong>of</strong><br />

the 850–500-hPa layer <strong>to</strong> 575 hPa.<br />

(850–575 hPa lapse rate, 5 WCD, LCL, and freez<strong>in</strong>g<br />

level). A cold front was located between the locations<br />

<strong>of</strong> the 2025 and 2031 UTC dropsondes. There was also<br />

a dryl<strong>in</strong>e <strong>in</strong>tersect<strong>in</strong>g the cold front, form<strong>in</strong>g a triple<br />

po<strong>in</strong>t <strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> the Learjet flight track (not<br />

shown). The dryl<strong>in</strong>e ran between the 2031 and 2034<br />

UTC dropsondes <strong>in</strong> Fig. 15, and this is where CI occurred<br />

as determ<strong>in</strong>ed from satellite imagery (not<br />

shown).<br />

NCAPE rapidly <strong>in</strong>creased <strong>in</strong> the vic<strong>in</strong>ity <strong>of</strong> the<br />

dryl<strong>in</strong>e, where convection <strong>in</strong>itiated, and peaked at<br />

about 0.11 m s 2 just 10–40 km eastward <strong>in</strong><strong>to</strong> the<br />

warmer and moister air (Fig. 15a). Although low-level<br />

shear was stronger rearward (i.e., westward) <strong>of</strong> the<br />

dryl<strong>in</strong>e and cold front, a relative maxima <strong>in</strong> the 0–3 km<br />

AGL shear (17.5 m s 1 ) was located just east <strong>of</strong> the<br />

dryl<strong>in</strong>e at 110 km (Fig. 15a). As a result, the develop<strong>in</strong>g<br />

convection on 24 May experienced <strong>in</strong>itially <strong>in</strong>creas<strong>in</strong>g<br />

and near-peak values <strong>of</strong> NCAPE and low-level shear.<br />

The CG flash percentage also rapidly <strong>in</strong>creased east-


1344 MONTHLY WEATHER REVIEW VOLUME 135<br />

FIG. 15. (a) NCAPE, low-level (0–3 km AGL) shear and (b) WCD, LCL, freez<strong>in</strong>g level, and 850–575-hPa lapse rate calculated from<br />

the 24 May 2002 dropsondes as a function <strong>of</strong> distance along the Learjet flight track. LCL and freez<strong>in</strong>g level heights are AGL. The time<br />

<strong>of</strong> each dropsonde release is <strong>in</strong>dicated and annotated by the dashed gray l<strong>in</strong>es <strong>in</strong> the vertical. The number <strong>of</strong> CG and CG flashes<br />

and the percent <strong>of</strong> positive polarity for s<strong>to</strong>rms with<strong>in</strong> each orange rectangular box <strong>in</strong> Fig. 14 are shown. The thick vertical dash–dotted<br />

l<strong>in</strong>e shows the location <strong>of</strong> CI for deep convection along the Learjet flight track as determ<strong>in</strong>ed by satellite imagery.


APRIL 2007 C A R E Y A N D B U F F A L O 1345<br />

ward <strong>of</strong> the dryl<strong>in</strong>e <strong>to</strong> a maximum value <strong>of</strong> 73% at a<br />

po<strong>in</strong>t centered just 35 km eastward (i.e., at 135 km<br />

along the Learjet flight track), apparently <strong>in</strong> response<br />

<strong>to</strong> these elevated values <strong>of</strong> NCAPE and low-level<br />

shear. Mov<strong>in</strong>g farther eastward, the NCAPE decreased<br />

dramatically (up <strong>to</strong> 36%) and the low-level shear<br />

dropped slightly. At the same time, the CG percentage<br />

also decreased significantly from the peak <strong>of</strong> 73%<br />

<strong>to</strong> only 24%, which is just below the subjective threshold<br />

required for positive s<strong>to</strong>rm status. The height <strong>of</strong> the<br />

LCL peaked just 20 km westward <strong>of</strong> the dryl<strong>in</strong>e (Fig.<br />

15b). Initial convection on 24 May was associated with<br />

an LCL between the maximum <strong>of</strong> 2200 and 1600 m<br />

AGL, which was measured just 10 km eastward <strong>of</strong> the<br />

dryl<strong>in</strong>e where convection <strong>in</strong>itiated (i.e., 110 km along<br />

the Learjet flight track). The LCL cont<strong>in</strong>ued <strong>to</strong> decrease<br />

eastward, reach<strong>in</strong>g 1000 m AGL at the last dropsonde<br />

located at 217 km along the Learjet flight track<br />

(i.e., about 120 km eastward <strong>of</strong> where convection <strong>in</strong>itiated).<br />

The height <strong>of</strong> the freez<strong>in</strong>g level gradually <strong>in</strong>creased<br />

eastward along the flight track. Comb<strong>in</strong><strong>in</strong>g the<br />

LCL and freez<strong>in</strong>g level heights on 24 May <strong>in</strong> Fig. 15b,<br />

the WCD <strong>in</strong>creased noticeably eastward <strong>of</strong> the dryl<strong>in</strong>e<br />

from a value that was between 1300 and 2000 m where<br />

convection <strong>in</strong>itiated <strong>to</strong> just above 2700 m about 120 km<br />

eastward <strong>of</strong> the po<strong>in</strong>t <strong>of</strong> CI. The 850–575-hPa lapse<br />

rates peaked at 8.4°C km 1 just west <strong>of</strong> where convection<br />

<strong>in</strong>itiated, and steadily decreased <strong>to</strong> the east, reach<strong>in</strong>g<br />

a value <strong>of</strong> 6.8°C km 1 at the last dropsonde location,<br />

about 120 km eastward <strong>of</strong> CI (Fig. 15b). The dramatic<br />

decrease <strong>in</strong> the CG fraction eastward <strong>of</strong> the<br />

dryl<strong>in</strong>e on 24 May was accompanied by a slight <strong>in</strong>crease<br />

<strong>in</strong> the freez<strong>in</strong>g level, a significant lower<strong>in</strong>g <strong>of</strong> the LCL,<br />

an associated noteworthy <strong>in</strong>crease <strong>in</strong> the WCD, and a<br />

modest decrease <strong>in</strong> the 850–575-hPa lapse rate. These<br />

trends are consistent with the relationships found between<br />

CG flash polarity and these environmental parameters<br />

<strong>in</strong> the grouped and regional comparisons.<br />

5. Summary and discussion<br />

The analysis <strong>of</strong> proximity sound<strong>in</strong>g data associated<br />

with both positive and negative CG dom<strong>in</strong>ant severe<br />

s<strong>to</strong>rms dur<strong>in</strong>g IHOP_2002 over the central United<br />

States clearly demonstrated significant and systematic<br />

differences <strong>in</strong> their mesoscale environments. When<br />

compared <strong>to</strong> negative s<strong>to</strong>rms, positive s<strong>to</strong>rms occurred<br />

<strong>in</strong> environments associated with a drier low- <strong>to</strong> midlevel<br />

troposphere (i.e., lower surface dewpo<strong>in</strong>t, mean mix<strong>in</strong>g<br />

ratio <strong>in</strong> the lowest 100 hPa, and PW from the surface <strong>to</strong><br />

400 hPa), higher CBH (i.e., higher LCL), smaller (i.e.,<br />

shallower) WCD, stronger conditional <strong>in</strong>stability (i.e.,<br />

larger 850–500- and 700–500-hPa lapse rates), larger<br />

0–3 km AGL w<strong>in</strong>d shear, stronger 0–2 km AGL s<strong>to</strong>rmrelative<br />

w<strong>in</strong>d speed, larger CAPE/NCAPE <strong>in</strong> the<br />

mixed-phase zone (i.e., LFC <strong>to</strong> 40°C NCAPE, and<br />

larger 10° <strong>to</strong> 40°C CAPE). Differences <strong>in</strong> the WCD<br />

<strong>of</strong> positive and negative s<strong>to</strong>rms were by far the most<br />

dramatic, suggest<strong>in</strong>g an important role for this parameter<br />

<strong>in</strong> controll<strong>in</strong>g CG lightn<strong>in</strong>g polarity.<br />

These results support the hypothesis that broader,<br />

stronger updrafts and larger supercooled liquid water<br />

contents <strong>in</strong> the mixed-phase zone <strong>of</strong> convection cause<br />

the positive charg<strong>in</strong>g <strong>of</strong> graupel and hail via NIC (e.g.,<br />

Saunders and Peck 1998) and the subsequent formation<br />

<strong>of</strong> a midlevel (i.e., 10° <strong>to</strong> 20°C) positive charge region<br />

and enhanced production <strong>of</strong> CG lightn<strong>in</strong>g as has<br />

been recently observed (Rust and MacGorman 2002;<br />

Lang et al. 2004; Wiens et al. 2005). Stronger updrafts <strong>in</strong><br />

positive s<strong>to</strong>rms were apparently generated thermodynamically<br />

from larger and more efficiently utilized conditional<br />

<strong>in</strong>stability/buoyancy <strong>in</strong> the mixed-phase zone<br />

and dynamically from stronger low-level shear and<br />

s<strong>to</strong>rm relative <strong>in</strong>flow, similar <strong>to</strong> the results <strong>of</strong> GW02.<br />

Higher CBH <strong>in</strong> positive s<strong>to</strong>rms likely resulted <strong>in</strong><br />

broader updrafts and reduced entra<strong>in</strong>ment, allow<strong>in</strong>g<br />

more <strong>of</strong> the potentially available buoyancy (i.e., CAPE<br />

and NCAPE) <strong>to</strong> be realized and effectively caus<strong>in</strong>g<br />

stronger updrafts as was first suggested by W05. Additionally,<br />

the lower PW content <strong>in</strong> positive s<strong>to</strong>rms was<br />

shown <strong>to</strong> be generally associated with stronger peak<br />

updrafts <strong>in</strong> simulated convection by McCaul et al.<br />

(2005) due <strong>to</strong> reduced water load<strong>in</strong>g, among other effects.<br />

The broader and stronger updrafts apparently<br />

generated larger supercooled liquid water contents <strong>in</strong><br />

the mixed-phase zone <strong>of</strong> positive s<strong>to</strong>rms by suppress<strong>in</strong>g<br />

coalescence and reduc<strong>in</strong>g dry air entra<strong>in</strong>ment, thus expla<strong>in</strong><strong>in</strong>g<br />

how environmental conditions can systematically<br />

control CG lightn<strong>in</strong>g polarity. Dramatically reduced<br />

WCD <strong>in</strong> positive s<strong>to</strong>rms apparently also <strong>in</strong>creased<br />

the supercooled liquid water content by<br />

drastically reduc<strong>in</strong>g the ra<strong>in</strong>out <strong>of</strong> available cloud water<br />

via collision–coalescence as was postulated by W05.<br />

Although McCaul and Cohen (2002) found that LFC<br />

and not the LCL controlled overturn<strong>in</strong>g efficiency and<br />

the degree <strong>of</strong> entra<strong>in</strong>ment <strong>in</strong> simulated deep convection,<br />

we found no statistically significant difference between<br />

the LFCs <strong>of</strong> positive and negative s<strong>to</strong>rm environments<br />

(Table 8). The mean LFC <strong>in</strong> both environments<br />

was just above the stated optimal range for<br />

overturn<strong>in</strong>g efficiency <strong>in</strong> McCaul and Cohen (2002). In<br />

their study, the strongest updraft case actually resulted<br />

from high LFC and low LCL (i.e., large LFC–LCL<br />

depth) because <strong>of</strong> the comb<strong>in</strong>ed <strong>in</strong>fluence <strong>of</strong> <strong>in</strong>creased<br />

overturn<strong>in</strong>g efficiency (high LFC) and the elim<strong>in</strong>ation<br />

<strong>of</strong> outflow dom<strong>in</strong>ance (low LCL). In our study, the<br />

LFC–LCL depth was significantly larger <strong>in</strong> the negative


1346 MONTHLY WEATHER REVIEW VOLUME 135<br />

(mean <strong>of</strong> 1.6 km) s<strong>to</strong>rms as compared <strong>to</strong> the positive<br />

s<strong>to</strong>rms (mean <strong>of</strong> 0.7 km; Table 5). Of course, this result<br />

was primarily associated with the much lower LCL <strong>in</strong><br />

negative s<strong>to</strong>rms s<strong>in</strong>ce the LFC was comparable. All else<br />

be<strong>in</strong>g equal, this result would suggest that negative<br />

s<strong>to</strong>rms have more efficient overturn<strong>in</strong>g, stronger updrafts,<br />

and less dilution <strong>of</strong> cloud water and buoyancy<br />

accord<strong>in</strong>g <strong>to</strong> McCaul and Cohen (2002), which would<br />

reject our hypothesis. It is also possible that other fac<strong>to</strong>rs<br />

(e.g., larger NCAPE and 0–3-km shear <strong>in</strong> positive<br />

s<strong>to</strong>rms) masked this effect such that positive s<strong>to</strong>rms still<br />

had stronger, broader updrafts and higher supercooled<br />

liquid water contents. In McCaul and Cohen (2002) the<br />

LCL and LFC were equivalent <strong>in</strong> most simulations and<br />

the LFC was used as a proxy for the moist layer depth<br />

such that all levels beneath the LFC were characterized<br />

by equally enhanced e and hence a most unstable CAPE<br />

(MUCAPE) layer. The exam<strong>in</strong>ed sound<strong>in</strong>gs <strong>in</strong> our study<br />

did not share these idealized characteristics as there<br />

was typically CIN (i.e., LCL LFC) and the LFC was<br />

on average 200 hPa above the lifted parcel level (LPL)<br />

<strong>of</strong> the most unstable parcel <strong>in</strong> the lowest 300 hPa from<br />

which MUCAPE was calculated. S<strong>in</strong>ce the LFC was<br />

therefore not an adequate proxy for moist layer depth<br />

<strong>in</strong> our sample, the McCaul and Cohen (2002) conclusions<br />

regard<strong>in</strong>g the role <strong>of</strong> LFC on overturn<strong>in</strong>g efficiency<br />

may not apply <strong>to</strong> our study. S<strong>in</strong>ce our sample was small<br />

(48 sound<strong>in</strong>gs), a future study us<strong>in</strong>g a larger dataset <strong>of</strong><br />

real sound<strong>in</strong>gs is required <strong>to</strong> confirm this suggestion.<br />

a. Large supercooled water contents <strong>in</strong> high based<br />

severe s<strong>to</strong>rms: An apparent paradox<br />

In our hypothesized scenario there is an apparent<br />

paradox <strong>in</strong> associat<strong>in</strong>g larger supercooled liquid water<br />

contents with the higher CBH <strong>of</strong> positive s<strong>to</strong>rms. Because<br />

<strong>of</strong> their lower and warmer cloud bases and more<br />

moist boundary layers (higher low-level mix<strong>in</strong>g ratio<br />

and dewpo<strong>in</strong>t), negative s<strong>to</strong>rms have higher adiabatic<br />

liquid water and hence the potential for higher actual<br />

cloud water contents than positive s<strong>to</strong>rms, especially if<br />

the cores are nearly undiluted as one might expect <strong>in</strong><br />

supercell convection. 6 Hence, <strong>to</strong> argue that high cloud<br />

6 Of course, it is common sense that this is not where the bulk<br />

<strong>of</strong> the electrification is occurr<strong>in</strong>g s<strong>in</strong>ce there are not sufficient<br />

number concentrations <strong>of</strong> precipitation ice particles <strong>in</strong> the core<br />

updraft or weak echo region (WER) <strong>of</strong> a supercell. Verification <strong>of</strong><br />

this common sense idea is the “lightn<strong>in</strong>g hole” or absence <strong>of</strong><br />

lightn<strong>in</strong>g and <strong>in</strong>ferred significant charge <strong>in</strong> VHF-based lightn<strong>in</strong>g<br />

observations with<strong>in</strong> the supercell WER (e.g., Krehbiel et al. 2000).<br />

S<strong>in</strong>ce we are unsure <strong>of</strong> where the electrification is tak<strong>in</strong>g place<br />

relative <strong>to</strong> the draft structure, it would not be safe <strong>to</strong> assume that<br />

the parcel is undiluted where NIC is operative.<br />

base (small WCD) positive s<strong>to</strong>rms have more supercooled<br />

(T 0°C) cloud water than negative s<strong>to</strong>rms, we<br />

must somehow compensate for more adiabatic condensate<br />

be<strong>in</strong>g available <strong>in</strong> negative s<strong>to</strong>rms <strong>in</strong> the first place.<br />

We (as have others) make the follow<strong>in</strong>g arguments that<br />

might compensate for more available moisture <strong>in</strong> negative<br />

s<strong>to</strong>rms:<br />

1) High CBH (or LCL) <strong>in</strong> positive s<strong>to</strong>rms may translate<br />

<strong>in</strong><strong>to</strong> broader updrafts with less entra<strong>in</strong>ment<br />

(e.g., McCarthy 1974) s<strong>in</strong>ce the convective bubble<br />

scales like the CBH or cumulus-<strong>to</strong>pped boundary<br />

layer depth (e.g., Kaimal et al. 1976). Similar suggestions<br />

have been made by Michaud (1996, 1998),<br />

Lucas et al. (1994, 1996), Williams and Stanfill<br />

(2002), Zipser (2003), and W05. Less entra<strong>in</strong>ment<br />

would result <strong>in</strong> less dilution <strong>of</strong> cloud water and<br />

buoyancy <strong>in</strong> positive s<strong>to</strong>rms. Hence, there would<br />

also be more efficient process<strong>in</strong>g <strong>of</strong> CAPE and<br />

stronger updrafts <strong>in</strong> positive s<strong>to</strong>rms, all else be<strong>in</strong>g<br />

equal.<br />

2) Lower (higher) cloud bases <strong>in</strong> negative (positive)<br />

s<strong>to</strong>rms means greater (lesser) depth through which<br />

mix<strong>in</strong>g <strong>of</strong> environmental air can reduce cloud buoyancy<br />

and water as po<strong>in</strong>ted out by Michaud (1996,<br />

1998) and discussed by Lucas et al. (1996).<br />

3) In general CAPE between positive and negative<br />

s<strong>to</strong>rms was very similar <strong>in</strong> our study. However,<br />

CAPE between 10° and 40°C and NCAPE <strong>in</strong><br />

nearly all layers was stronger <strong>in</strong> positive s<strong>to</strong>rms provid<strong>in</strong>g<br />

stronger vertical accelerations <strong>in</strong> the mixedphase<br />

zone for NIC. As also found <strong>in</strong> GW02, lowlevel<br />

(0–3 km) shear was larger and hence dynamic<br />

pressure forces may be stronger <strong>in</strong> positive s<strong>to</strong>rms<br />

(e.g., Weisman and Klemp 1984). In comb<strong>in</strong>ation<br />

with potentially more efficient process<strong>in</strong>g <strong>of</strong> available<br />

CAPE (see above), the updraft <strong>in</strong> positive<br />

s<strong>to</strong>rms may be stronger. Postulated stronger updrafts<br />

<strong>in</strong> positive s<strong>to</strong>rms would have the effect <strong>of</strong><br />

suppress<strong>in</strong>g precipitation and <strong>in</strong>creas<strong>in</strong>g the cloud/<br />

precipitation fraction as was recently suggested by<br />

W05 <strong>in</strong> a similar context and as has long been known<br />

<strong>in</strong> the severe s<strong>to</strong>rm community (e.g., the echo-free<br />

vault; Brown<strong>in</strong>g 1964, 1965).<br />

4) Lower cloud base (and <strong>to</strong> a lesser extent higher<br />

freez<strong>in</strong>g height) <strong>in</strong> negative s<strong>to</strong>rms results <strong>in</strong> significantly<br />

larger WCD. Larger WCD would tend <strong>to</strong> <strong>in</strong>crease<br />

the efficiency <strong>of</strong> warm ra<strong>in</strong>–collision–<br />

coalescence processes (e.g., Rosenfeld and Woodley<br />

2003). Increased coalescence results <strong>in</strong> lower<strong>in</strong>g <strong>of</strong><br />

the cloud/precipitation fraction. Once precipitation<br />

is formed, this condensate is no longer available <strong>to</strong><br />

be l<strong>of</strong>ted <strong>in</strong><strong>to</strong> the mixed-phase zone as cloud water


APRIL 2007 C A R E Y A N D B U F F A L O 1347<br />

where it can affect s<strong>to</strong>rm electrification (Williams et<br />

al. 2002; W05). Enhanced condensate <strong>in</strong> the warm<br />

portion <strong>of</strong> negative s<strong>to</strong>rms can also <strong>in</strong>crease water<br />

load<strong>in</strong>g and frictional drag and reduce the updraft<br />

velocity before the ra<strong>in</strong> falls out, which feeds back<br />

<strong>in</strong><strong>to</strong> the po<strong>in</strong>ts above.<br />

5) Although less conv<strong>in</strong>c<strong>in</strong>g from the evidence presented,<br />

there was an observed tendency for more<br />

supercell characteristics <strong>in</strong> positive over negative<br />

s<strong>to</strong>rms (Table 3). From radar, supercell and multicell<br />

characteristics are evident <strong>in</strong> both positive and<br />

negative s<strong>to</strong>rms. However, <strong>in</strong>spection <strong>of</strong> loops <strong>of</strong><br />

low-level radar reflectivity from <strong>in</strong>dividual and regional<br />

composite Weather Surveillance Radar–1988<br />

Doppler (WSR-88D) imagery suggested that supercells<br />

were more common <strong>in</strong> positive s<strong>to</strong>rms. Limited<br />

analysis <strong>of</strong> WSR-88D Doppler velocity us<strong>in</strong>g the<br />

Warn<strong>in</strong>g Decision Support System–Integrated Information<br />

(WDSS-II) s<strong>of</strong>tware revealed significantly<br />

higher mesocyclone frequency and <strong>in</strong>tensity<br />

for the 15 June positive s<strong>to</strong>rms as compared <strong>to</strong> the<br />

15 June negative s<strong>to</strong>rms (not shown). This f<strong>in</strong>d<strong>in</strong>g<br />

suggests that supercell characteristics and hence dynamical<br />

forc<strong>in</strong>g were stronger and more common <strong>in</strong><br />

positive s<strong>to</strong>rms, at least on this particular day. This<br />

result is generally consistent with the environmental<br />

data (e.g., lower BRN, larger 0–3-km shear, larger<br />

0–3-km s<strong>to</strong>rm relative helicity, larger EHI, and<br />

roughly equivalent CAPE and 0–6-km shear), which<br />

encourage pressure perturbation dynamics associated<br />

with quasi-steady (supercell) forc<strong>in</strong>g <strong>of</strong> the updraft<br />

<strong>in</strong> positive s<strong>to</strong>rms. This dynamical forc<strong>in</strong>g<br />

alone could have 1) <strong>in</strong>creased updraft strength and<br />

2) decreased detrimental entra<strong>in</strong>ment <strong>of</strong> buoyancy<br />

and cloud water <strong>in</strong> positive s<strong>to</strong>rms, thus feed<strong>in</strong>g<br />

back on several issues above.<br />

The above suggestions are merely hypotheses based<br />

on plausible physical mechanisms and not observed<br />

facts. Nonetheless, the presented environmental data is<br />

consistent with these hypotheses. In the meantime, they<br />

wait further test<strong>in</strong>g with <strong>in</strong> situ observations and numerical<br />

cloud models <strong>to</strong> evaluate whether these compensat<strong>in</strong>g<br />

fac<strong>to</strong>rs <strong>in</strong>crease (decrease) the (cloud water)/<br />

(precipitation water) and the (actual supercooled water<br />

content)/(adiabatic supercooled water content) fractions<br />

<strong>in</strong> positive (negative) s<strong>to</strong>rms sufficiently <strong>to</strong> result<br />

<strong>in</strong> larger absolute supercooled cloud water contents <strong>in</strong><br />

positive s<strong>to</strong>rms. To understand how much <strong>of</strong> a compensat<strong>in</strong>g<br />

effect might be necessary, it is worthwhile <strong>to</strong><br />

consider differences <strong>in</strong> the mean adiabatic liquid water<br />

content for the exam<strong>in</strong>ed positive and negative s<strong>to</strong>rms.<br />

The mean adiabatic liquid water contents at 0°, 20°,<br />

and 40°C for negative s<strong>to</strong>rms were 6.0, 10.5, and 12.9<br />

gkg 1 , respectively, while for positive s<strong>to</strong>rms they were<br />

3.3, 7.0, and 9.8 g kg 1 , respectively. Note that the difference<br />

decreases significantly from the bot<strong>to</strong>m (0°C)<br />

<strong>to</strong> the <strong>to</strong>p (40°C) <strong>of</strong> the mixed-phase zone. In the<br />

middle <strong>of</strong> the mixed-phase zone (20°C) where NIC is<br />

the most effective, the mean adiabatic liquid water content<br />

<strong>in</strong> positive s<strong>to</strong>rms is about two-thirds <strong>of</strong> the magnitude<br />

<strong>in</strong> negative s<strong>to</strong>rms, provid<strong>in</strong>g some estimate <strong>of</strong><br />

the required compensat<strong>in</strong>g effects. Also note that the<br />

results confirm that potential water load<strong>in</strong>g effects <strong>in</strong><br />

the warm portion <strong>of</strong> the cloud are significantly higher <strong>in</strong><br />

negative s<strong>to</strong>rms, as expected.<br />

b. The relative roles <strong>of</strong> the various environmental<br />

fac<strong>to</strong>rs <strong>in</strong> controll<strong>in</strong>g CG lightn<strong>in</strong>g polarity<br />

It is important <strong>to</strong> note that not all <strong>of</strong> the environmental<br />

fac<strong>to</strong>rs summarized above were always equal <strong>in</strong><br />

importance for expla<strong>in</strong><strong>in</strong>g why each <strong>in</strong>dividual positive<br />

(negative) s<strong>to</strong>rm was apparently associated with stronger/broader<br />

(weaker/narrower) updrafts and larger<br />

(smaller) liquid water contents <strong>in</strong> the mixed-phase<br />

zone. Sometimes one environmental fac<strong>to</strong>r seem<strong>in</strong>gly<br />

compensated for another (Table 9 and Figs. 12 and 13),<br />

as would be expected, s<strong>in</strong>ce updraft forc<strong>in</strong>g and the<br />

production <strong>of</strong> large supercooled liquid water contents<br />

can come from any comb<strong>in</strong>ation <strong>of</strong> the mechanisms discussed<br />

above. Nonetheless, based on the observational<br />

evidence from IHOP_2002, the low-level moisture,<br />

LCL, and WCD must be highlighted as likely the most<br />

important environmental fac<strong>to</strong>rs for determ<strong>in</strong><strong>in</strong>g the<br />

dom<strong>in</strong>ant CG polarity <strong>in</strong> severe s<strong>to</strong>rms. Furthermore,<br />

the crucial role <strong>of</strong> the WCD may help expla<strong>in</strong> the apparent<br />

paradox that most severe s<strong>to</strong>rms actually produce<br />

predom<strong>in</strong>antly CG lightn<strong>in</strong>g (CRP03) despite<br />

also likely be<strong>in</strong>g associated with vigorous vertical<br />

drafts. As po<strong>in</strong>ted out by W05, the clima<strong>to</strong>logical overlap<br />

<strong>of</strong> shallow WCD (i.e., high cloud base) and large<br />

<strong>in</strong>stability likely def<strong>in</strong>es the geographic distribution <strong>of</strong><br />

positive s<strong>to</strong>rms shown <strong>in</strong> CRP03 and repeated <strong>in</strong> Fig. 1.<br />

Based on our results, we generally agree with this view<br />

but suggest that the occurrence <strong>of</strong> strong 0–3 kmAGL<br />

w<strong>in</strong>d shear be added <strong>to</strong> this union <strong>of</strong> environmental<br />

fac<strong>to</strong>rs, s<strong>in</strong>ce dynamic can equal or exceed thermodynamic<br />

forc<strong>in</strong>g <strong>of</strong> the updraft with<strong>in</strong> severe supercell<br />

s<strong>to</strong>rms over the central United States. In summary, sufficient<br />

moisture, <strong>in</strong>stability, and deep-layer shear <strong>in</strong><br />

comb<strong>in</strong>ation with enhanced NCAPE at heights below<br />

40°C, enhanced dynamical forc<strong>in</strong>g as seen <strong>in</strong> enhanced<br />

low-level shear and SREH, and a more efficient<br />

process<strong>in</strong>g <strong>of</strong> CAPE and moisture up <strong>to</strong> and <strong>in</strong> the<br />

supercooled mixed-phase zone associated with higher<br />

LCL and smaller WCD is the likely explanation for


1348 MONTHLY WEATHER REVIEW VOLUME 135<br />

7 Recall that there was no significant difference between the<br />

mean positive and negative region e and CAPE.<br />

positive s<strong>to</strong>rm occurrence <strong>in</strong> the central United States.<br />

As such, it should be clear that LCL or WCD alone<br />

cannot cause the conditions favorable for strong, broad<br />

updrafts and high supercooled water contents. For example,<br />

<strong>in</strong> the dry conditions <strong>of</strong> the desert Southwest<br />

CBH is very high and WCD very shallow or perhaps<br />

zero but yet positive s<strong>to</strong>rms are not common there.<br />

Clearly, this is because the other necessary conditions<br />

such as sufficient moisture and CAPE are not met. It is<br />

<strong>in</strong>terest<strong>in</strong>g <strong>to</strong> note <strong>in</strong> Fig. 1 and CRP03 that a few localized<br />

areas <strong>in</strong> the desert Southwest and Intermounta<strong>in</strong><br />

West experience a large fraction <strong>of</strong> positive s<strong>to</strong>rms.<br />

The sample size <strong>of</strong> severe s<strong>to</strong>rm reports there is small<br />

so the result may not be significant. Nonetheless, it<br />

would be <strong>in</strong>terest<strong>in</strong>g <strong>to</strong> conduct a similar study <strong>in</strong> those<br />

areas under the proper conditions <strong>to</strong> see if our results<br />

are confirmed.<br />

It is important <strong>to</strong> note that several <strong>of</strong> the analyzed<br />

environmental parameters are highly correlated (i.e., or<br />

not <strong>in</strong>dependent such as surface dewpo<strong>in</strong>t temperature,<br />

low-level mix<strong>in</strong>g ratio, LCL, and WCD). Hence it could<br />

be difficult <strong>to</strong> assign causality <strong>to</strong> any one parameter<br />

<strong>in</strong>dividually. In a future study, it would be worthwhile<br />

<strong>to</strong> construct a multiple-l<strong>in</strong>ear regression model and<br />

analysis <strong>of</strong> variance <strong>to</strong> determ<strong>in</strong>e the relative importance<br />

<strong>of</strong> several parameters simultaneously and <strong>to</strong><br />

elim<strong>in</strong>ate redundant variables. Furthermore, other potential<br />

hypotheses expla<strong>in</strong><strong>in</strong>g CG lightn<strong>in</strong>g polarity <strong>in</strong>dependent<br />

<strong>of</strong> the LCL and WCD but related <strong>in</strong> some<br />

other fashion <strong>to</strong> the role <strong>of</strong> low-level moisture on cloud<br />

electrification and lightn<strong>in</strong>g, which are coupled closely<br />

<strong>to</strong> mixed-phase cloud microphysical processes, are possible.<br />

As po<strong>in</strong>ted out by CRP03, it is not entirely clear from<br />

Smith et al. (2000) why s<strong>to</strong>rms should be positive dom<strong>in</strong>ant<br />

<strong>in</strong> the e gradient region. We suggest that the e<br />

gradient region (on the western and northern sides) <strong>of</strong><br />

the ridge <strong>in</strong> moist entropy is a preferred area for positive<br />

s<strong>to</strong>rms (Smith et al. 2000; CRP03) not simply because<br />

<strong>of</strong> a difference <strong>in</strong> e or CAPE values, 7 but rather<br />

because the comb<strong>in</strong>ation <strong>of</strong> environmental parameters<br />

we found <strong>to</strong> be favorable for CG production are<br />

likely <strong>to</strong> be found <strong>in</strong> unison there, <strong>in</strong>clud<strong>in</strong>g sufficient<br />

low-level moisture, ample <strong>in</strong>stability that tends <strong>to</strong> be<br />

more concentrated <strong>in</strong> the supercooled mixed-phase<br />

zone, high LCL, shallow WCD, and large 0–3-km shear.<br />

It is important <strong>to</strong> note that this comb<strong>in</strong>ation <strong>of</strong> variables<br />

is consistent with the possibility <strong>of</strong> stronger updrafts<br />

and supercooled liquid water contents <strong>in</strong> positive<br />

s<strong>to</strong>rms, as required by the <strong>in</strong>verted-dipole hypothesis.<br />

c. The potential role <strong>of</strong> aerosols<br />

Because <strong>of</strong> <strong>in</strong>sufficient aerosol data, we have not explored<br />

the potential effect <strong>of</strong> cloud condensation nuclei<br />

(CCN) on cloud electrification and CG lightn<strong>in</strong>g polarity.<br />

Accord<strong>in</strong>g <strong>to</strong> W05, CCN enhancements <strong>in</strong> the<br />

boundary layer could lead <strong>to</strong> a reduction <strong>in</strong> droplet size,<br />

the suppression <strong>of</strong> coalescence (e.g., Williams et al.<br />

2002; Rosenfeld and Woodley 2003), a result<strong>in</strong>g <strong>in</strong>crease<br />

<strong>in</strong> the supercooled liquid water content, and<br />

hence the production <strong>of</strong> enhanced CG lightn<strong>in</strong>g as<br />

discussed above. This mechanism may expla<strong>in</strong> the reported<br />

relationship between anomalously high aerosol<br />

concentrations associated with biomass burn<strong>in</strong>g and the<br />

occurrence <strong>of</strong> positive s<strong>to</strong>rms over the southern pla<strong>in</strong>s<br />

dur<strong>in</strong>g the spr<strong>in</strong>g <strong>of</strong> 1998 (Lyons et al. 1998; Murray et<br />

al. 2000). As po<strong>in</strong>ted out by W05 and shown by Lang<br />

and Rutledge (2006), the roles <strong>of</strong> extraord<strong>in</strong>ary thermodynamic,<br />

dynamic, and CCN conditions are <strong>of</strong>ten<br />

difficult <strong>to</strong> dist<strong>in</strong>guish s<strong>in</strong>ce they <strong>of</strong>ten occur simultaneously.<br />

Nonetheless, Lang and Rutledge (2006) show<br />

that CG lightn<strong>in</strong>g enhancements downw<strong>in</strong>d <strong>of</strong> a forest<br />

fire <strong>in</strong> Colorado were more consistent with a causative<br />

role for elevated CBH and reduced WCD than for<br />

smoke aerosols. To confirm our results and <strong>to</strong> evaluate<br />

the potential impact <strong>of</strong> CCN on CG lightn<strong>in</strong>g polarity,<br />

we suggest a special focus be placed on measur<strong>in</strong>g the<br />

simultaneous thermodynamic, w<strong>in</strong>d, and aerosol properties<br />

<strong>of</strong> <strong>in</strong>flow air <strong>in</strong><strong>to</strong> deep convection, <strong>in</strong>clud<strong>in</strong>g severe<br />

s<strong>to</strong>rms, <strong>in</strong> a variety <strong>of</strong> conditions.<br />

d. <strong>Environmental</strong> control <strong>of</strong> supercell type and CG<br />

polarity: A comparison <strong>of</strong> results<br />

There are three supercells types <strong>in</strong>clud<strong>in</strong>g LP (e.g.,<br />

Blueste<strong>in</strong> and Parks 1983), classic (e.g., Brown<strong>in</strong>g<br />

1964), and high precipitation (HP; e.g., Doswell and<br />

Burgess 1993) that have been identified and def<strong>in</strong>ed by<br />

the amount <strong>of</strong> precipitation they produce and where<br />

the precipitation is deposited relative <strong>to</strong> their respective<br />

updrafts and mesocyclones (e.g., Doswell and Burgess<br />

1993; Rasmussen and Straka 1998). The distribution<br />

and amount <strong>of</strong> precipitation <strong>in</strong> supercells is thought <strong>to</strong><br />

have important implications for s<strong>to</strong>rm microphysics<br />

that could affect supercell dynamics, <strong>in</strong>clud<strong>in</strong>g <strong>to</strong>rnadogenesis<br />

(e.g., Brooks et al. 1994a) and cloud electrification<br />

(e.g., MacGorman and Burgess 1994).<br />

Early observations showed LP supercell s<strong>to</strong>rms <strong>to</strong> be<br />

CG dom<strong>in</strong>ant (Curran and Rust 1992; Branick and<br />

Doswell 1992) and HP supercells <strong>to</strong> be CG dom<strong>in</strong>ant


APRIL 2007 C A R E Y A N D B U F F A L O 1349<br />

(Branick and Doswell 1992), fuel<strong>in</strong>g speculation that a<br />

potentially exclusive relationship between supercell<br />

type and CG lightn<strong>in</strong>g polarity might exist. However,<br />

Blueste<strong>in</strong> and MacGorman (1998) present observations<br />

<strong>of</strong> an LP supercell that was dom<strong>in</strong>ated by CG lightn<strong>in</strong>g.<br />

They also mention unpublished observations <strong>of</strong><br />

two HP supercells dom<strong>in</strong>ated by CG lightn<strong>in</strong>g. Classic<br />

supercells appear <strong>to</strong> be primarily CG dom<strong>in</strong>ant<br />

but can also be associated with CG dom<strong>in</strong>ant behavior<br />

(MacGorman and Nielsen 1991; MacGorman and<br />

Burgess 1994). The HP supercells are typically associated<br />

with dom<strong>in</strong>ant CG lightn<strong>in</strong>g (e.g., Branick and<br />

Doswell 1992; MacGorman and Burgess 1994; Knupp<br />

et al. 2003). Transitions from LP <strong>to</strong> classic (and classic<br />

<strong>to</strong> HP) supercell types are <strong>of</strong>ten associated with correspond<strong>in</strong>g<br />

transitions from positive <strong>to</strong> negative CGdom<strong>in</strong>ant<br />

lightn<strong>in</strong>g behavior (Seimon 1993; MacGorman<br />

and Burgess 1994). Dom<strong>in</strong>ant negative CG supercells<br />

that transition from classic <strong>to</strong> HP type have also<br />

been associated with <strong>in</strong>creas<strong>in</strong>g CG lightn<strong>in</strong>g production<br />

(Knupp et al. 2003). Multicellular s<strong>to</strong>rms (sometimes<br />

with embedded supercells) have been observed <strong>to</strong><br />

produce primarily CG lightn<strong>in</strong>g (MacGorman and<br />

Burgess 1994; Carey and Rutledge 1998) even though<br />

most multicell s<strong>to</strong>rms produce primarily CG lightn<strong>in</strong>g.<br />

In summary, although there is a general tendency<br />

for LP (HP) supercells <strong>to</strong> be positive (negative) CG<br />

dom<strong>in</strong>ant, there are exceptions <strong>to</strong> this tendency as<br />

noted by Blueste<strong>in</strong> and MacGorman (1998).<br />

There are two studies on the relationship between<br />

environmental conditions and supercell type (LP, classic,<br />

HP) that are relevant <strong>to</strong> our discussion: Blueste<strong>in</strong><br />

and Parks (1983) and Rasmussen and Straka (1998).<br />

Blueste<strong>in</strong> and Parks (1983) found that the LCL was<br />

significantly higher <strong>in</strong> LP (1.8 km AGL) than classic<br />

(1.4 km AGL) supercells. Mean mix<strong>in</strong>g ratio <strong>in</strong> the lowest<br />

kilometer was correspond<strong>in</strong>gly lower <strong>in</strong> LP (11.9 g<br />

kg 1 ) versus classic (13.5 g kg 1 ) supercells. The PW<br />

was lower <strong>in</strong> LP (2.8 cm) than classic (3.3 cm) supercells.<br />

All <strong>of</strong> these results are consistent with the fact<br />

that LP supercells are most common near the dryl<strong>in</strong>e as<br />

were two <strong>of</strong> the four positive s<strong>to</strong>rms regions <strong>in</strong> this<br />

study (23 and 24 May). On the other hand, Rasmussen<br />

and Straka (1998) did not f<strong>in</strong>d statistically significant<br />

differences <strong>in</strong> the LCL or low-level mix<strong>in</strong>g ratio between<br />

the three supercell types. They did f<strong>in</strong>d that PW<br />

was higher <strong>in</strong> HP versus LP and classic supercells. Rasmussen<br />

and Straka (1998) also found that upper-level<br />

s<strong>to</strong>rm relative flow at 9–10 km is stronger <strong>in</strong> LP s<strong>to</strong>rms<br />

and comparatively weak <strong>in</strong> HP s<strong>to</strong>rms with classic supercells<br />

fall<strong>in</strong>g <strong>in</strong> between. They speculate that precipitation<br />

efficiency is relatively lowered (raised) <strong>in</strong> LP<br />

(HP) s<strong>to</strong>rms due <strong>to</strong> the decreased (<strong>in</strong>creased) recirculation<br />

<strong>of</strong> hydrometeors <strong>in</strong><strong>to</strong> the supercell updraft associated<br />

with the stronger (weaker) anvil-level w<strong>in</strong>ds.<br />

Given the general tendency for LP (classic) supercells<br />

<strong>to</strong> be associated with positive (negative) s<strong>to</strong>rms,<br />

our results are <strong>in</strong> good agreement with the Blueste<strong>in</strong><br />

and Parks (1983) f<strong>in</strong>d<strong>in</strong>gs regard<strong>in</strong>g the meteorological<br />

environments <strong>of</strong> LP and classic supercells. A comparison<br />

with Rasmussen and Straka (1998) is less encourag<strong>in</strong>g<br />

as only the PW results appear <strong>to</strong> be consistent<br />

with our study (i.e., high PW is associated with HP<br />

supercells and negative s<strong>to</strong>rms as expected). As a matter<br />

<strong>of</strong> fact, we <strong>in</strong>spected upper-level s<strong>to</strong>rm relative w<strong>in</strong>d<br />

speed based on their results and the known tendencies<br />

between supercell type and CG lightn<strong>in</strong>g polarity. We<br />

found no statistically significant difference between the<br />

9–11-km s<strong>to</strong>rm relative w<strong>in</strong>d speed <strong>of</strong> positive and<br />

negative s<strong>to</strong>rms dur<strong>in</strong>g IHOP_2002. Reasons for the<br />

discourag<strong>in</strong>g comparison with Rasmussen and Straka<br />

(1998) are unknown but they noted that their sound<strong>in</strong>gs<br />

may not have adequately sampled the low-level moisture<br />

conditions <strong>of</strong> the supercells. We also note that<br />

there are known exceptions regard<strong>in</strong>g the general tendencies<br />

<strong>of</strong> CG polarity with supercell type so any comparison<br />

between our study and supercell environment<br />

studies must be viewed with some caution.<br />

Nonetheless, the fairly consistent relationship between<br />

supercell type and CG polarity can apparently be<br />

expla<strong>in</strong>ed by at least one common l<strong>in</strong>k: precipitation<br />

efficiency. By def<strong>in</strong>ition, LP supercells are extremely<br />

precipitation <strong>in</strong>efficient, imply<strong>in</strong>g that a relatively large<br />

fraction <strong>of</strong> available condensate rema<strong>in</strong>s <strong>in</strong> cloud form<br />

(both water and ice) and is not converted <strong>to</strong> precipitation.<br />

As suggested earlier, the large implied (cloud water)/(precipitation<br />

water) fraction <strong>in</strong> LP supercells<br />

could allow for the positive NIC <strong>of</strong> precipitation ice, the<br />

generation <strong>of</strong> a midlevel positive charge layer, and positive<br />

lightn<strong>in</strong>g. Based on our results and Blueste<strong>in</strong> and<br />

Parks (1983), we hypothesize that high CBH and shallow<br />

WCD <strong>in</strong> LP s<strong>to</strong>rms is likely an important causal<br />

fac<strong>to</strong>r for the precipitation <strong>in</strong>efficiency, associated LP<br />

structure, and CG lightn<strong>in</strong>g, although other fac<strong>to</strong>rs<br />

likely play a role. The HP supercell is by def<strong>in</strong>ition a<br />

relatively efficient converter <strong>of</strong> cloud water <strong>to</strong> precipitation<br />

and hence would be characterized by a relatively<br />

smaller (cloud water)/(precipitation water) fraction.<br />

Although <strong>in</strong>consistent with Rasmussen and Straka<br />

(1998), we suggest that relatively low CBH and deep<br />

WCD might be preferentially associated with HP supercells<br />

and hence could be an important causal mechanism<br />

for their precipitation efficiency and CG lightn<strong>in</strong>g<br />

production. Of course, other physical fac<strong>to</strong>rs, such


1350 MONTHLY WEATHER REVIEW VOLUME 135<br />

as those highlighted by Rasmussen and Straka (1998),<br />

may also be important for both phenomena.<br />

e. Tornadoes, CG lightn<strong>in</strong>g polarity reversals, and<br />

the LCL<br />

As discussed earlier, CG lightn<strong>in</strong>g polarity reversals<br />

have been documented <strong>in</strong> some supercell s<strong>to</strong>rms (e.g.,<br />

Seimon 1993; MacGorman and Burgess 1994; Perez et<br />

al. 1997; Smith et al. 2000). The polarity reversals documented<br />

<strong>in</strong> these studies are typically from dom<strong>in</strong>ant<br />

positive <strong>to</strong> dom<strong>in</strong>ant negative CG lightn<strong>in</strong>g. Although<br />

not common, an abrupt polarity shift from mostly positive<br />

<strong>to</strong> mostly negative CG lightn<strong>in</strong>g is sometimes associated<br />

with <strong>to</strong>rnadogenesis and a significant (F2 or<br />

greater) <strong>to</strong>rnado on the ground (Seimon 1993;<br />

MacGorman and Burgess 1994; Perez et al. 1997). Although<br />

the potential causative fac<strong>to</strong>rs for both significant<br />

<strong>to</strong>rnadoes and CG lightn<strong>in</strong>g polarity are many,<br />

complex, and the subject <strong>of</strong> current debate, our study <strong>in</strong><br />

comb<strong>in</strong>ation with recent studies on the environmental<br />

conditions associated with significant <strong>to</strong>rnadic supercells<br />

may help expla<strong>in</strong> the occasional co<strong>in</strong>cidence between<br />

<strong>to</strong>rnadogenesis and CG lightn<strong>in</strong>g polarity reversals.<br />

Among other fac<strong>to</strong>rs, Rasmussen and Blanchard<br />

(1998) found that LCL was one <strong>of</strong> the best environmental<br />

discrim<strong>in</strong>a<strong>to</strong>rs between supercells that produce<br />

significant <strong>to</strong>rnadoes (<strong>to</strong>rnadic) and those that do not<br />

(non<strong>to</strong>rnadic). An LCL 800 m was associated with<br />

significant <strong>to</strong>rnadoes and 1200 m was associated with<br />

decreas<strong>in</strong>g likelihood <strong>of</strong> significant <strong>to</strong>rnadoes. More recent<br />

studies (Craven et al. 2002b; Thompson et al. 2003;<br />

Rasmussen 2003) have confirmed the strong correlation<br />

between significant <strong>to</strong>rnadoes <strong>in</strong> supercells and low<br />

LCL, although Rasmussen (2003) raises some concerns,<br />

which are associated with his methodology for isolat<strong>in</strong>g<br />

supercells, regard<strong>in</strong>g the causality <strong>of</strong> the result. S<strong>in</strong>ce<br />

our study shows that low LCL is highly correlated with<br />

dom<strong>in</strong>ant negative CG lightn<strong>in</strong>g, it is possible that sudden<br />

CG polarity shifts from dom<strong>in</strong>ant positive <strong>to</strong> dom<strong>in</strong>ant<br />

negative are associated with a rapid decrease <strong>in</strong><br />

the LCL (<strong>in</strong>crease <strong>in</strong> the low-level humidity) and hence<br />

associated with an <strong>in</strong>creased probability <strong>of</strong> a significant<br />

<strong>to</strong>rnado. Clearly, future studies should cont<strong>in</strong>ue <strong>to</strong> focus<br />

on the correlation between low-level moisture and<br />

supercell type, <strong>to</strong>rnado potential, and CG lightn<strong>in</strong>g polarity<br />

with an emphasis on observ<strong>in</strong>g and model<strong>in</strong>g the<br />

physical and dynamical fac<strong>to</strong>rs that could confirm or<br />

reject causal relationships. Based on the strong spatial<br />

correlation between high (<strong>in</strong>tracloud) IC–CG lightn<strong>in</strong>g<br />

ratio and positive CG fraction <strong>in</strong> the central United<br />

States (Boccippio et al. 2001), we would also add <strong>to</strong>tal<br />

lightn<strong>in</strong>g flash rate and the IC–CG lightn<strong>in</strong>g ratio <strong>to</strong> the<br />

list <strong>of</strong> s<strong>to</strong>rm properties that may be at least partially<br />

controlled by low-level humidity or CBH, as did W05.<br />

f. Reject<strong>in</strong>g the tilted-dipole hypothesis<br />

Table 8 provides conv<strong>in</strong>c<strong>in</strong>g evidence reject<strong>in</strong>g the<br />

tilted-dipole hypothesis (e.g., Brook et al. 1982) for expla<strong>in</strong><strong>in</strong>g<br />

dom<strong>in</strong>ant positive CG lightn<strong>in</strong>g <strong>in</strong> severe convection.<br />

The lack <strong>of</strong> a significant difference <strong>of</strong> the deeplayer<br />

(0–6 km) shear and s<strong>to</strong>rm-relative w<strong>in</strong>d speeds <strong>in</strong><br />

the mid- <strong>to</strong> upper troposphere (4–6, 6–10, 9–11 km<br />

AGL; EL) between positive and negative s<strong>to</strong>rms makes<br />

it difficult <strong>to</strong> imag<strong>in</strong>e how a positive s<strong>to</strong>rm could tilt the<br />

thunders<strong>to</strong>rm dipole more than a negative s<strong>to</strong>rm. Although<br />

some early studies seem <strong>to</strong> favor a role for the<br />

tilted-dipole hypothesis <strong>in</strong> supercells (e.g., Curran and<br />

Rust 1992; Branick and Doswell 1992) and mesoscale<br />

convective systems (MCSs; e.g., Orville et al. 1988;<br />

S<strong>to</strong>lzenburg 1990), early evidence reject<strong>in</strong>g this hypothesis<br />

can be found <strong>in</strong> Reap and MacGorman (1989), who<br />

found no clear relationship between deep-layer shear<br />

and CG polarity us<strong>in</strong>g observational and modelgenerated<br />

environmental data. Even Curran and Rust<br />

(1992) could only claim that strong deep-layer shear<br />

was likely a necessary yet not sufficient condition for<br />

the production <strong>of</strong> CG lightn<strong>in</strong>g. Recent <strong>in</strong> situ/<br />

balloon E-field and VHF-based 3D <strong>to</strong>tal lightn<strong>in</strong>g observations<br />

<strong>in</strong> supercells (e.g., Rust and MacGorman<br />

2002; MacGorman et al. 2005; Wiens et al. 2005) have<br />

<strong>in</strong>ferred vertical charge pr<strong>of</strong>iles that support the presence<br />

<strong>of</strong> an enhanced midlevel positive charge (<strong>in</strong> a generally<br />

<strong>in</strong>verted charge structure) as the most common<br />

source for positive CG flashes <strong>in</strong> many High Pla<strong>in</strong>s supercells<br />

that produce dom<strong>in</strong>ant positive CG lightn<strong>in</strong>g<br />

and not a tilted upper-level positive charge source as<br />

orig<strong>in</strong>ally hypothesized for w<strong>in</strong>ter s<strong>to</strong>rms by Brook et<br />

al. (1982). Us<strong>in</strong>g radar, NLDN, and VHF-based 3D <strong>to</strong>tal<br />

lightn<strong>in</strong>g observations, Carey et al. (2005) demonstrated<br />

that the bulk <strong>of</strong> the CG lightn<strong>in</strong>g production<br />

and <strong>in</strong>ferred charge structure <strong>in</strong> an MCS was <strong>in</strong>consistent<br />

with the basic premise <strong>of</strong> the tilted dipole hypothesis.<br />

Results here<strong>in</strong> merely add <strong>to</strong> the accumulat<strong>in</strong>g evidence<br />

from earlier environmental and more recent field<br />

campaign studies that reject the tilted-dipole hypothesis.<br />

Acknowledgments. We wish <strong>to</strong> acknowledge fruitful<br />

discussions with Earle Williams, Walter Petersen, Kenneth<br />

Cumm<strong>in</strong>s, Daniel Rosenfeld, Timothy Lang, and<br />

Donald MacGorman. We are grateful <strong>to</strong> Earle Williams<br />

for provid<strong>in</strong>g early access <strong>to</strong> ongo<strong>in</strong>g research.<br />

Earle Williams, Ted Mansell, and an anonymous reviewer<br />

provided meticulous and helpful reviews that<br />

improved the manuscript. The NLDN lightn<strong>in</strong>g data


APRIL 2007 C A R E Y A N D B U F F A L O 1351<br />

were obta<strong>in</strong>ed from Vaisala, Inc., under the direction <strong>of</strong><br />

Jerry Guynes <strong>of</strong> TAMU. We thank Brandon Ely and<br />

Scott Steiger for assistance with IDL code. This research<br />

was supported by National Science Foundation<br />

Grants ATM-0233780 and ATM-0442011.<br />

APPENDIX<br />

Selection <strong>of</strong> Proximity Sound<strong>in</strong>g Data<br />

As described <strong>in</strong> detail by Brooks et al. (1994b), obta<strong>in</strong><strong>in</strong>g<br />

proximity sound<strong>in</strong>gs that are representative <strong>of</strong><br />

the meteorological conditions experienced by convection<br />

is not a trivial task. The goal <strong>in</strong> select<strong>in</strong>g proximity<br />

sound<strong>in</strong>gs is <strong>to</strong> select only those sound<strong>in</strong>gs that sampled<br />

the <strong>in</strong>flow air <strong>of</strong> the s<strong>to</strong>rm(s) <strong>of</strong> <strong>in</strong>terest. However, spatial<br />

and temporal variability with<strong>in</strong> the environments <strong>of</strong><br />

severe s<strong>to</strong>rms is the rule, rather than the exception, and<br />

it is thus <strong>of</strong>ten difficult <strong>to</strong> obta<strong>in</strong> representative sound<strong>in</strong>gs.<br />

Fortunately, the multitude <strong>of</strong> sound<strong>in</strong>gs launched<br />

dur<strong>in</strong>g IHOP_2002 mitigated this problem.<br />

Three issues need <strong>to</strong> be considered and accounted for<br />

when select<strong>in</strong>g representative proximity sound<strong>in</strong>gs: 1)<br />

spatial variability <strong>of</strong> environmental conditions, 2) temporal<br />

variability <strong>of</strong> environmental conditions, and 3)<br />

the sampl<strong>in</strong>g by sound<strong>in</strong>gs <strong>of</strong> conditions that are not<br />

representative <strong>of</strong> the <strong>in</strong>flow air <strong>of</strong> the s<strong>to</strong>rm(s) <strong>of</strong> <strong>in</strong>terest,<br />

due <strong>to</strong> fac<strong>to</strong>rs such as convective contam<strong>in</strong>ation<br />

and the presence <strong>of</strong> boundaries (e.g., fronts, dryl<strong>in</strong>es,<br />

outflow boundaries; Brooks et al. 1994b). To account<br />

for the first and second issues, distance and time constra<strong>in</strong>ts<br />

<strong>of</strong> 100 km and 3 h, respectively, were used. To<br />

account for the third issue, all sound<strong>in</strong>gs satisfy<strong>in</strong>g the<br />

first and second criteria were <strong>in</strong>dividually <strong>in</strong>spected by<br />

hand for signatures <strong>of</strong> convective contam<strong>in</strong>ation (e.g.,<br />

the lower troposphere cooled and stabilized by outflow,<br />

the upper troposphere moistened by anvils, the w<strong>in</strong>d<br />

structure altered dramatically) and for signatures that<br />

the sound<strong>in</strong>g sampled a different air mass than that <strong>in</strong><br />

which the s<strong>to</strong>rms <strong>of</strong> <strong>in</strong>terest developed (e.g., the sound<strong>in</strong>g<br />

was launched on the opposite side <strong>of</strong> a front,<br />

dryl<strong>in</strong>e, or outflow boundary from where the convection<br />

developed) us<strong>in</strong>g surface observations, radar reflectivity<br />

imagery, visible satellite imagery, and ground<br />

flash plots. S<strong>in</strong>ce the proximity sound<strong>in</strong>g dataset compiled<br />

strongly dictates the results <strong>of</strong> a study, great detail<br />

and care were taken <strong>in</strong> assembl<strong>in</strong>g this dataset. From<br />

hundreds <strong>of</strong> sound<strong>in</strong>gs launched on the 6 days <strong>in</strong>vestigated<br />

<strong>in</strong> this study, 48 sound<strong>in</strong>gs were chosen as representative<br />

<strong>in</strong>flow proximity sound<strong>in</strong>gs. Half <strong>of</strong> these (24)<br />

represent the mesoscale environments <strong>of</strong> positive<br />

s<strong>to</strong>rms and the other half (24) represent the environments<br />

<strong>of</strong> negative s<strong>to</strong>rms (Table 2).<br />

In compil<strong>in</strong>g a proximity sound<strong>in</strong>g dataset, compet<strong>in</strong>g<br />

forces exist between assembl<strong>in</strong>g a large dataset by<br />

enforc<strong>in</strong>g less str<strong>in</strong>gent requirements on the sound<strong>in</strong>gs<br />

used, and assembl<strong>in</strong>g a dataset truly representative <strong>of</strong><br />

s<strong>to</strong>rm <strong>in</strong>flow air by enforc<strong>in</strong>g more str<strong>in</strong>gent requirements<br />

on the sound<strong>in</strong>gs used. Naturally, the latter approach<br />

results <strong>in</strong> a smaller proximity sound<strong>in</strong>g dataset<br />

(Brooks et al. 1994b). This study placed greater emphasis<br />

on us<strong>in</strong>g only those proximity sound<strong>in</strong>gs truly representative<br />

<strong>of</strong> <strong>in</strong>flow air than on the assemblage <strong>of</strong> a<br />

large dataset. Even so, the size <strong>of</strong> our proximity sound<strong>in</strong>g<br />

dataset is sufficiently large <strong>to</strong> produce statistically<br />

significant and robust results. Nonetheless, future studies<br />

should strive <strong>to</strong> compile a larger sample <strong>of</strong> <strong>in</strong>flow<br />

proximity sound<strong>in</strong>gs associated with positive and negative<br />

s<strong>to</strong>rms <strong>in</strong> order <strong>to</strong> confirm the f<strong>in</strong>d<strong>in</strong>gs here<strong>in</strong>.<br />

REFERENCES<br />

Blanchard, D. O., 1998: Assess<strong>in</strong>g the vertical distribution <strong>of</strong> convective<br />

available potential energy. Wea. Forecast<strong>in</strong>g, 13, 870–<br />

877.<br />

Blueste<strong>in</strong>, H. B., and C. R. Parks, 1983: A synoptic and pho<strong>to</strong>graphic<br />

clima<strong>to</strong>logy <strong>of</strong> low-precipitation severe thunders<strong>to</strong>rms<br />

<strong>in</strong> the southern pla<strong>in</strong>s. Mon. Wea. Rev., 111, 2034–<br />

2046.<br />

——, and D. R. MacGorman, 1998: Evolution <strong>of</strong> cloud-<strong>to</strong>-ground<br />

lightn<strong>in</strong>g characteristics and s<strong>to</strong>rm structure <strong>in</strong> the Spearman,<br />

Texas, <strong>to</strong>rnadic supercells <strong>of</strong> 31 May 1990. Mon. Wea. Rev.,<br />

126, 1451–1467.<br />

Boccippio, D. J., K. L. Cumm<strong>in</strong>s, H. J. Christian, and S. J. Goodman,<br />

2001: Comb<strong>in</strong>ed satellite- and surface-based estimation<br />

<strong>of</strong> the <strong>in</strong>tracloud–cloud-<strong>to</strong>-ground lightn<strong>in</strong>g ratio over the<br />

cont<strong>in</strong>ental United States. Mon. Wea. Rev., 129, 108–122.<br />

Bol<strong>to</strong>n, D., 1980: The computation <strong>of</strong> equivalent potential temperature.<br />

Mon. Wea. Rev., 108, 1046–1053.<br />

Branick, M. L., and C. A. Doswell III, 1992: An observation <strong>of</strong> the<br />

relationship between supercell structure and lightn<strong>in</strong>g<br />

ground-strike polarity. Wea. Forecast<strong>in</strong>g, 7, 143–149.<br />

Brook, M., M. Nakano, P. Krehbiel, and T. Takeuti, 1982: The<br />

electrical structure <strong>of</strong> the Hokuriku w<strong>in</strong>ter thunders<strong>to</strong>rms. J.<br />

Geophys. Res., 87, 1207–1215.<br />

Brooks, H. B., C. A. Doswell III, and R. B. Wilhelmson, 1994a:<br />

The role <strong>of</strong> midtropospheric w<strong>in</strong>ds <strong>in</strong> the evolution and ma<strong>in</strong>tenance<br />

<strong>of</strong> low-level mesocyclones. Mon. Wea. Rev., 122,<br />

126–136.<br />

——, ——, and J. Cooper, 1994b: On the environments <strong>of</strong> <strong>to</strong>rnadic<br />

and non<strong>to</strong>rnadic mesocyclones. Wea. Forecast<strong>in</strong>g, 9,<br />

606–618.<br />

Brown<strong>in</strong>g, K. A., 1964: Airflow and precipitation trajec<strong>to</strong>ries<br />

with<strong>in</strong> severe local s<strong>to</strong>rms which travel <strong>to</strong> the right <strong>of</strong> the<br />

w<strong>in</strong>ds. J. Atmos. Sci., 21, 634–639.<br />

——, 1965: Some <strong>in</strong>ferences about the updraft with<strong>in</strong> a severe<br />

local s<strong>to</strong>rm. J. Atmos. Sci., 22, 669–677.<br />

Carey, L. D., and S. A. Rutledge, 1998: Electrical and multiparameter<br />

radar observations <strong>of</strong> a severe hails<strong>to</strong>rm. J. Geophys.<br />

Res., 103, 13 979–14 000.<br />

——, and ——, 2003: Characteristics <strong>of</strong> cloud-<strong>to</strong>-ground lightn<strong>in</strong>g<br />

<strong>in</strong> severe and nonsevere s<strong>to</strong>rms over the central United


1352 MONTHLY WEATHER REVIEW VOLUME 135<br />

States from 1989–1998. J. Geophys. Res., 108, 4483,<br />

doi:10.1029/2002JD002951.<br />

——, W. A. Petersen, and S. A. Rutledge, 2003a: Evolution <strong>of</strong><br />

cloud-<strong>to</strong>-ground lightn<strong>in</strong>g and s<strong>to</strong>rm structure <strong>in</strong> the Spencer,<br />

SD, supercell <strong>of</strong> 30 May 1998. Mon. Wea. Rev., 131, 1811–<br />

1831.<br />

——, S. A. Rutledge, and W. A. Petersen, 2003b: The relationship<br />

between severe s<strong>to</strong>rm reports and cloud-<strong>to</strong>-ground lightn<strong>in</strong>g<br />

polarity <strong>in</strong> the contiguous United States from 1989–98. Mon.<br />

Wea. Rev., 131, 1211–1228.<br />

——, M. J. Murphy, T. L. McCormick, and N. W. S. Demetriades,<br />

2005: <strong>Lightn<strong>in</strong>g</strong> location relative <strong>to</strong> s<strong>to</strong>rm structure <strong>in</strong> a lead<strong>in</strong>g-l<strong>in</strong>e,<br />

trail<strong>in</strong>g-stratiform mesoscale convective system. J.<br />

Geophys. Res., 110, D03105, doi:10.1029/2003JD004371.<br />

Craven, J. P., R. E. Jewell, and H. E. Brooks, 2002a: Comparison<br />

between observed convective cloud-base heights and lift<strong>in</strong>g<br />

condensation level for two different lifted parcels. Wea. Forecast<strong>in</strong>g,<br />

17, 885–890.<br />

——, H. E. Brooks, and J. A. Hart, 2002b: Basel<strong>in</strong>e clima<strong>to</strong>logy <strong>of</strong><br />

sound<strong>in</strong>g derived parameters associated with deep, moist<br />

convection. Prepr<strong>in</strong>ts, 21st Conf. on Severe Local S<strong>to</strong>rms, San<br />

An<strong>to</strong>nio, TX, Amer. Meteor. Soc., 643–646.<br />

Cumm<strong>in</strong>s, K. L., M. J. Murphy, E. A. Bardo, W. L. Hiscox, R. B.<br />

Pyle, and A. E. Pifer, 1998: A comb<strong>in</strong>ed TOA/MDF technology<br />

upgrade <strong>of</strong> the U.S. National <strong>Lightn<strong>in</strong>g</strong> Detection Network.<br />

J. Geophys. Res., 103, 9035–9044.<br />

——, J. A. Cramer, C. J. Biagi, E. P. Krider, J. Jerauld, M. Uman,<br />

and V. Rakov, 2006: The U.S. National <strong>Lightn<strong>in</strong>g</strong> Detection<br />

Network: Post-upgrade status. Prepr<strong>in</strong>ts, Second Conf. on the<br />

Meteorological Applications <strong>of</strong> <strong>Lightn<strong>in</strong>g</strong> Data, Atlanta, GA,<br />

Amer. Meteor. Soc., CD-ROM, P6.1.<br />

Curran, E. B., and W. D. Rust, 1992: Positive ground flashes produced<br />

by low-precipitation thunders<strong>to</strong>rms <strong>in</strong> Oklahoma on<br />

26 April 1984. Mon. Wea. Rev., 120, 544–553.<br />

Davies-Jones, R. P., D. Burgess, and M. Foster, 1990: Test <strong>of</strong><br />

helicity as a <strong>to</strong>rnado forecast parameter. Prepr<strong>in</strong>ts, 16th Conf.<br />

on Severe Local S<strong>to</strong>rms, Kananaskis Park, AB, Canada,<br />

Amer. Meteor. Soc., 588–592.<br />

Doswell, C. A., III, and D. W. Burgess, 1993: Tornadoes and <strong>to</strong>rnadic<br />

s<strong>to</strong>rms: A review <strong>of</strong> conceptual models. The Tornado:<br />

Its Structure, Dynamics, Prediction, and Hazards, Geophys.<br />

Monogr., Vol. 28, Amer. Geophys. Union, 161–172.<br />

——, and E. N. Rasmussen, 1994: The effect <strong>of</strong> neglect<strong>in</strong>g the<br />

virtual temperature correction on CAPE calculations. Wea.<br />

Forecast<strong>in</strong>g, 9, 619–623.<br />

Gilmore, M. S., and L. J. Wicker, 2002: Influences <strong>of</strong> the local<br />

environment on supercell cloud-<strong>to</strong>-ground lightn<strong>in</strong>g, radar<br />

characteristics, and severe weather on 2 June 1995. Mon.<br />

Wea. Rev., 130, 2349–2372.<br />

Hart, J. A., and W. D. Korotky, 1991: The SHARP workstation<br />

v1. 50 user’s guide. National Weather Service, NOAA, 30 pp.<br />

[Available from NWS Eastern Region Headquarters, Scientific<br />

Services Division, 630 Johnson Ave., Bohemia, NY<br />

11716.]<br />

Heckert, N. A., and J. J. Filliben, 2003: NIST Handbook 148:<br />

DATAPLOT reference manual, Volume I: Commands. National<br />

Institute <strong>of</strong> Standards and Technology Handbook Series,<br />

June 2003. [Available onl<strong>in</strong>e at http://www.itl.nist.gov/<br />

div898/s<strong>of</strong>tware/dataplot/refman1/homepage.htm.]<br />

Houze, R. A., Jr., 1993: <strong>Cloud</strong> Dynamics. Academic Press, 573 pp.<br />

Jayaratne, E. R., C. P. R. Saunders, and J. Hallett, 1983: Labora<strong>to</strong>ry<br />

studies <strong>of</strong> the charg<strong>in</strong>g <strong>of</strong> s<strong>of</strong>t-hail dur<strong>in</strong>g ice crystal<br />

<strong>in</strong>teractions. Quart. J. Roy. Meteor. Soc., 109, 609–630.<br />

Jorgensen, D. P., and M. A. LeMone, 1989: Vertical velocity characteristics<br />

<strong>of</strong> oceanic convection. J. Atmos. Sci., 46, 621–640.<br />

Kaimal, J. C., J. C. Wyngaard, D. A. Haugen, O. R. Cote, Y.<br />

Izumi, S. J. Caughey, and C. J. Read<strong>in</strong>gs, 1976: Turbulence<br />

structures <strong>in</strong> the convective boundary layer. J. Atmos. Sci., 33,<br />

2152–2169.<br />

Knapp, D. I., 1994: Us<strong>in</strong>g cloud-<strong>to</strong>-ground lightn<strong>in</strong>g data <strong>to</strong> identify<br />

<strong>to</strong>rnadic thunders<strong>to</strong>rm signatures and nowcast severe<br />

weather. Natl. Wea. Digest, 19, 35–42.<br />

Knight, C. A., and N. C. Knight, 2001: Hails<strong>to</strong>rms. Severe Convective<br />

S<strong>to</strong>rms, Meteor. Monogr., No. 50, Amer. Meteor. Soc.,<br />

223–248.<br />

Knupp, K. R., S. Peach, and S. Goodman, 2003: Variations <strong>in</strong><br />

cloud-<strong>to</strong>-ground lightn<strong>in</strong>g characteristics among three adjacent<br />

<strong>to</strong>rnadic supercell s<strong>to</strong>rms over the Tennessee valley region.<br />

Mon. Wea. Rev., 131, 172–188.<br />

Krehbiel, P. R., R. J. Thomas, W. Rison, T. Haml<strong>in</strong>, J. Harl<strong>in</strong>, and<br />

M. Davis, 2000: GPS-based mapp<strong>in</strong>g system reveals lightn<strong>in</strong>g<br />

<strong>in</strong>side s<strong>to</strong>rms. Eos, Trans. Amer. Geophys. Union, 81, 21–25.<br />

Lang, T. J., and S. A. Rutledge, 2002: Relationships between convective<br />

s<strong>to</strong>rm k<strong>in</strong>ematics, precipitation, and lightn<strong>in</strong>g. Mon.<br />

Wea. Rev., 130, 2492–2506.<br />

——, and ——, 2006: <strong>Cloud</strong>-<strong>to</strong>-ground lightn<strong>in</strong>g downw<strong>in</strong>d <strong>of</strong> the<br />

2002 Hayman forest fire <strong>in</strong> Colorado. Geophys. Res. Lett., 33,<br />

L03804, doi:10.1029/2005GL024608.<br />

——, and Coauthors, 2004: The Severe Thunders<strong>to</strong>rm Electrification<br />

and Precipitation Study. Bull. Amer. Meteor. Soc., 85,<br />

1107–1126.<br />

Lucas, C., E. J. Zipser, and M. A. LeMone, 1994: Vertical velocity<br />

<strong>in</strong> oceanic convection <strong>of</strong>f tropical Australia. J. Atmos. Sci., 51,<br />

3183–3193.<br />

——, ——, and ——, 1996: Reply. J. Atmos. Sci., 53, 1212–1214.<br />

Lyons, W. A., T. E. Nelson, E. R. Williams, J. Cramer, and T.<br />

Turner, 1998: Enhanced positive cloud-<strong>to</strong>-ground lightn<strong>in</strong>g <strong>in</strong><br />

thunders<strong>to</strong>rms <strong>in</strong>gest<strong>in</strong>g smoke. Science, 282, 77–81.<br />

MacGorman, D. R., and K. E. Nielsen, 1991: <strong>Cloud</strong>-<strong>to</strong>-groundlightn<strong>in</strong>g<br />

<strong>in</strong> a <strong>to</strong>rnadic s<strong>to</strong>rm on 8 May 1986. Mon. Wea. Rev.,<br />

119, 1557–1574.<br />

——, and D. W. Burgess, 1994: Positive cloud-<strong>to</strong>-ground lightn<strong>in</strong>g<br />

<strong>in</strong> <strong>to</strong>rnadic s<strong>to</strong>rms and hails<strong>to</strong>rms. Mon. Wea. Rev., 122,<br />

1671–1697.<br />

——, W. D. Rust, P. Krehbiel, W. Rison, E. Brun<strong>in</strong>g, and K.<br />

Wiens, 2005: The electrical structure <strong>of</strong> two supercell s<strong>to</strong>rms<br />

dur<strong>in</strong>g STEPS. Mon. Wea. Rev., 133, 2583–2607.<br />

McCarthy, J., 1974: Field verification <strong>of</strong> the relationship between<br />

entra<strong>in</strong>ment rate and cumulus cloud diameter. J. Atmos. Sci.,<br />

31, 1028–1039.<br />

McCaul, E. W., Jr., and M. L. Weisman, 2001: The sensitivity <strong>of</strong><br />

simulated supercell structure and <strong>in</strong>tensity <strong>to</strong> variations <strong>in</strong> the<br />

shapes <strong>of</strong> environmental buoyancy and shear pr<strong>of</strong>iles. Mon.<br />

Wea. Rev., 129, 664–687.<br />

——, and C. Cohen, 2002: The impact on simulated s<strong>to</strong>rm structure<br />

and <strong>in</strong>tensity <strong>of</strong> variations <strong>in</strong> the mixed layer and moist<br />

layer depths. Mon. Wea. Rev., 130, 1722–1748.<br />

——, ——, and C. Kirkpatrick, 2005: The sensitivity <strong>of</strong> simulated<br />

s<strong>to</strong>rm structure, <strong>in</strong>tensity, and precipitation efficiency <strong>to</strong> environmental<br />

temperature. Mon. Wea. Rev., 133, 3015–3037.<br />

Michaud, L. M., 1996: Comments on “Convective available potential<br />

energy <strong>in</strong> the environment <strong>of</strong> oceanic and cont<strong>in</strong>ental<br />

clouds.” J. Atmos. Sci., 53, 1209–1211.<br />

——, 1998: Entra<strong>in</strong>ment and detra<strong>in</strong>ment required <strong>to</strong> expla<strong>in</strong> updraft<br />

properties and work dissipation. Tellus, 50A, 283–301.<br />

Moller, A. R., 2001: Severe local s<strong>to</strong>rm forecast<strong>in</strong>g. Severe Con-


APRIL 2007 C A R E Y A N D B U F F A L O 1353<br />

vective S<strong>to</strong>rms, Meteor. Monogr., No. 50, Amer. Meteor. Soc.,<br />

433–480.<br />

Moncrieff, M. W., and M. J. Miller, 1976: The dynamics and simulation<br />

<strong>of</strong> tropical cumulonimbus and squall l<strong>in</strong>es. Quart. J.<br />

Roy. Meteor. Soc., 102, 373–394.<br />

Murray, N. D., R. E. Orville, and G. R. Huff<strong>in</strong>es, 2000: Effect <strong>of</strong><br />

pollution from Central American fires on cloud-<strong>to</strong>-ground<br />

lightn<strong>in</strong>g <strong>in</strong> May 1998. Geophys. Res. Lett., 27, 2249–2252.<br />

Orville, R. E., and G. R. Huff<strong>in</strong>es, 2001: <strong>Cloud</strong>-<strong>to</strong>-ground lightn<strong>in</strong>g<br />

<strong>in</strong> the United States: NLDN results <strong>in</strong> the first decade,<br />

1989–98. Mon. Wea. Rev., 129, 1179–1193.<br />

——, R. W. Henderson, and L. F. Bosart, 1988: Bipole patterns<br />

revealed by lightn<strong>in</strong>g locations <strong>in</strong> mesoscale s<strong>to</strong>rm systems.<br />

Geophys. Res. Lett., 15, 129–132.<br />

Perez, A. H., L. J. Wicker, and R. E. Orville, 1997: Characteristics<br />

<strong>of</strong> cloud-<strong>to</strong>-ground lightn<strong>in</strong>g associated with violent <strong>to</strong>rnadoes.<br />

Wea. Forecast<strong>in</strong>g, 12, 428–437.<br />

Rasmussen, E. N., 2003: Ref<strong>in</strong>ed supercell and <strong>to</strong>rnado forecast<br />

parameters. Wea. Forecast<strong>in</strong>g, 18, 530–535.<br />

——, and D. O. Blanchard, 1998: A basel<strong>in</strong>e clima<strong>to</strong>logy <strong>of</strong><br />

sound<strong>in</strong>g-derived supercell and <strong>to</strong>rnado forecast parameters.<br />

Wea. Forecast<strong>in</strong>g, 13, 1148–1164.<br />

——, and J. M. Straka, 1998: Variations <strong>in</strong> supercell morphology.<br />

Part I: Observations <strong>of</strong> the role <strong>of</strong> upper-level s<strong>to</strong>rm-relative<br />

flow. Mon. Wea. Rev., 126, 2406–2421.<br />

Reap, R. M., and D. R. MacGorman, 1989: <strong>Cloud</strong>-<strong>to</strong>-ground lightn<strong>in</strong>g:<br />

Clima<strong>to</strong>logical characteristics and relationships <strong>to</strong><br />

model fields, radar observations, and severe local s<strong>to</strong>rms.<br />

Mon. Wea. Rev., 117, 518–535.<br />

Rosenfeld, D., and W. L. Woodley, 2003: Clos<strong>in</strong>g the 5-year circle:<br />

From cloud seed<strong>in</strong>g <strong>to</strong> space and back <strong>to</strong> climate change<br />

through precipitation physics. <strong>Cloud</strong> Systems, Hurricanes,<br />

and the Tropical Ra<strong>in</strong>fall Measur<strong>in</strong>g Mission (TRMM), Meteor.<br />

Monogr., No. 51, Amer. Meteor. Soc., 59–80.<br />

Rotunno, R., J. B. Klemp, and M. L. Weisman, 1988: A theory for<br />

strong, long-lived squall l<strong>in</strong>es. J. Atmos. Sci., 45, 463–485.<br />

Rust, W. D., and D. R. MacGorman, 2002: Possibly <strong>in</strong>vertedpolarity<br />

electrical structures <strong>in</strong> thunders<strong>to</strong>rms dur<strong>in</strong>g STEPS.<br />

Geophys. Res. Lett., 29, 1571, doi:10.1029/2001GL014303.<br />

Saunders, C. P. R., and S. L. Peck, 1998: Labora<strong>to</strong>ry studies <strong>of</strong> the<br />

<strong>in</strong>fluence <strong>of</strong> the rime accretion rate on charge transfer dur<strong>in</strong>g<br />

crystal/graupel collisions. J. Geophys. Res., 103, 13 949–<br />

13 956.<br />

——, W. D. Keith, and R. P. Mitzeva, 1991: The effect <strong>of</strong> liquid<br />

water on thunders<strong>to</strong>rm charg<strong>in</strong>g. J. Geophys. Res., 96,<br />

11 007–11 017.<br />

Seimon, A., 1993: Anomalous cloud-<strong>to</strong>-ground lightn<strong>in</strong>g <strong>in</strong> an F5-<br />

<strong>to</strong>rnado-produc<strong>in</strong>g supercell thunders<strong>to</strong>rm on 28 August<br />

1990. Bull. Amer. Meteor. Soc., 74, 189–203.<br />

Smith, S. B., J. G. LaDue, and D. R. MacGorman, 2000: The relationship<br />

between cloud-<strong>to</strong>-ground lightn<strong>in</strong>g polarity and<br />

surface equivalent potential temperature dur<strong>in</strong>g three <strong>to</strong>rnadic<br />

outbreaks. Mon. Wea. Rev., 128, 3320–3328.<br />

Stephens, M. A., 1974: EDF statistics for goodness <strong>of</strong> fit and some<br />

comparisons. J. Amer. Stat. Assoc., 69, 730–737.<br />

S<strong>to</strong>lzenburg, M., 1990: Characteristics <strong>of</strong> the bipolar pattern <strong>of</strong><br />

lightn<strong>in</strong>g locations observed <strong>in</strong> 1988 thunders<strong>to</strong>rms. Bull.<br />

Amer. Meteor. Soc., 71, 1331–1338.<br />

——, 1994: Observations <strong>of</strong> high ground flash densities <strong>of</strong> positive<br />

lightn<strong>in</strong>g <strong>in</strong> summertime thunders<strong>to</strong>rms. Mon. Wea. Rev.,<br />

122, 1740–1750.<br />

——, W. D. Rust, and T. C. Marshall, 1998: Electrical structure <strong>in</strong><br />

thunders<strong>to</strong>rm convective regions 3. Synthesis. J. Geophys.<br />

Res., 103, 14 097–14 108.<br />

Takahashi, T., 1978: Rim<strong>in</strong>g electrification as a charge generation<br />

mechanism <strong>in</strong> thunders<strong>to</strong>rms. J. Atmos. Sci., 35, 1536–1548.<br />

Thompson, R. L., R. Edwards, J. A. Hart, K. L. Elmore, and P.<br />

Markowski, 2003: Close proximity sound<strong>in</strong>gs with<strong>in</strong> supercell<br />

environments obta<strong>in</strong>ed from the Rapid Update Cycle. Wea.<br />

Forecast<strong>in</strong>g, 18, 1243–1261.<br />

Wacker, R. S., and R. E. Orville, 1999a: Changes <strong>in</strong> measured<br />

lightn<strong>in</strong>g flash count and return stroke peak current after the<br />

1994 U.S. National <strong>Lightn<strong>in</strong>g</strong> Detection Network upgrade, 1,<br />

Observations. J. Geophys. Res., 104, 2151–2158.<br />

——, and ——, 1999b: Changes <strong>in</strong> measured lightn<strong>in</strong>g flash count<br />

and return stroke peak current after the 1994 U.S. National<br />

<strong>Lightn<strong>in</strong>g</strong> Detection Network upgrade, 2, Theory. J. Geophys.<br />

Res., 104, 2159–2162.<br />

Weckwerth, T. M., and Coauthors, 2004: An overview <strong>of</strong> the International<br />

H 2 O Project (IHOP_2002) and some prelim<strong>in</strong>ary<br />

highlights. Bull. Amer. Meteor. Soc., 85, 253–277.<br />

Weisman, M. L., and J. B. Klemp, 1982: The dependence <strong>of</strong> numerically<br />

simulated convective s<strong>to</strong>rms on vertical w<strong>in</strong>d shear<br />

and buoyancy. Mon. Wea. Rev., 110, 504–520.<br />

——, and ——, 1984: The structure and classification <strong>of</strong> numerically<br />

simulated convective s<strong>to</strong>rms <strong>in</strong> directionally vary<strong>in</strong>g<br />

w<strong>in</strong>d shears. Mon. Wea. Rev., 112, 2479–2498.<br />

——, and ——, 1986: Characteristics <strong>of</strong> isolated convective<br />

s<strong>to</strong>rms. Mesoscale Meteorology and Forecast<strong>in</strong>g, P. S. Ray,<br />

Ed., Amer. Meteor. Soc. 331–358.<br />

——, and R. Rotunno, 2004: “A theory for strong long-lived<br />

squall l<strong>in</strong>es” revisited. J. Atmos. Sci., 61, 361–382.<br />

Wiens, K. C., S. A. Rutledge, and S. A. Tessendorf, 2005: The 29<br />

June 2000 supercell observed dur<strong>in</strong>g STEPS. Part II: <strong>Lightn<strong>in</strong>g</strong><br />

and charge structure. J. Atmos. Sci., 62, 4151–4177.<br />

Wilks, D. S., 1995: Statistical Methods <strong>in</strong> the Atmospheric Sciences.<br />

Academic Press, 467 pp.<br />

Williams, E. R., 1989: The tripole structure <strong>of</strong> thunders<strong>to</strong>rms. J.<br />

Geophys. Res., 94, 13 151–13 167.<br />

——, 2001: The electrification <strong>of</strong> severe s<strong>to</strong>rms. Severe Convective<br />

S<strong>to</strong>rms, Meteor. Monogr., No. 50, Amer. Meteor. Soc., 527–<br />

561.<br />

——, and N. Renno, 1993: An analysis <strong>of</strong> the conditional <strong>in</strong>stability<br />

<strong>of</strong> the tropical atmosphere. Mon. Wea. Rev., 121, 21–36.<br />

——, and S. Stanfill, 2002: The physical orig<strong>in</strong> <strong>of</strong> the land-ocean<br />

contrast <strong>in</strong> lightn<strong>in</strong>g activity. C. R. Phys., 3, 1277–1292.<br />

——, and Coauthors, 2002: Contrast<strong>in</strong>g convective regimes over<br />

the Amazon: Implications for cloud electrification. J. Geophys.<br />

Res., 107, 8082, doi:10.1029/2001JD000380.<br />

——, V. Mushtak, D. Rosenfeld, S. Goodman, and D. Boccippio,<br />

2005: Thermodynamic conditions favorable <strong>to</strong> superlative<br />

thunders<strong>to</strong>rm updraft, mixed phase microphysics and lightn<strong>in</strong>g<br />

flash rate. Atmos. Res., 76, 288–306.<br />

Wilson, C. T. R., 1920: Investigations on lightn<strong>in</strong>g discharges and<br />

on the electric field <strong>of</strong> thunders<strong>to</strong>rms. Philos. Trans. Roy.<br />

Soc. London, 221A, 73–115.<br />

Zipser, E. J., 2003: Some views on “Hot Towers” after 50 years <strong>of</strong><br />

tropical field programs and two years <strong>of</strong> TRMM data. <strong>Cloud</strong><br />

Systems, Hurricanes, and the Tropical Ra<strong>in</strong>fall Measur<strong>in</strong>g<br />

Mission (TRMM), Meteor. Monogr., No. 51, Amer. Meteor.<br />

Soc., 49–58.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!