• Atkins, N. T., R. M. Wakimoto, and C. L. Ziegler, 1998: Observations of the finescale structure of a dryline during VORTEX 95. Mon. Wea. Rev., 126 , 525550.

    • Search Google Scholar
    • Export Citation
  • Benjamin, J. B., 1968: Gravity current and related phenomena. J. Fluid Mech., 31 , 209248.

  • Charba, J., 1974: Application of gravity current model to analysis of squall-line gust front. Mon. Wea. Rev., 102 , 140156.

  • Cheung, T. K., and C. G. Little, 1990: Meteorological tower, microbarograph array, and sodar observations of solitary-like waves in the nocturnal boundary layer. J. Atmos. Sci., 47 , 25162536.

    • Search Google Scholar
    • Export Citation
  • Clarke, R. H., 1984: Colliding sea breezes and the creation of internal atmospheric bore waves: Two-dimensional numerical studies. Aust. Meteor. Mag., 32 , 207226.

    • Search Google Scholar
    • Export Citation
  • Clarke, R. H., R. K. Smith, and D. G. Reid, 1981: The morning glory of the Gulf of Carpentaria: An atmospheric undular bore. Mon. Wea. Rev., 109 , 17261750.

    • Search Google Scholar
    • Export Citation
  • Coleman, T. A., K. R. Knupp, and D. Herzmann, 2009: The spectacular undular bore in Iowa on 2 October 2007. Mon. Wea. Rev., 137 , 495503.

    • Search Google Scholar
    • Export Citation
  • Crook, N. A., 1986: The effect of ambient stratification and moisture on the motion of atmospheric undular bores. J. Atmos. Sci., 43 , 171181.

    • Search Google Scholar
    • Export Citation
  • Crook, N. A., 1988: Trapping of low-level internal gravity waves. J. Atmos. Sci., 45 , 15331541.

  • Crook, N. A., and M. J. Miller, 1985: A numerical and analytical study of atmospheric undular bores. Quart. J. Roy. Meteor. Soc., 111 , 225242.

    • Search Google Scholar
    • Export Citation
  • Davis, C., and Coauthors, 2004: The bow echo and MCV experiment: Observations and opportunities. Bull. Amer. Meteor. Soc., 85 , 10751093.

    • Search Google Scholar
    • Export Citation
  • Doviak, R. J., and R. Ge, 1984: An atmospheric solitary gust observed with a Doppler radar, a tall tower and a surface network. J. Atmos. Sci., 41 , 25592573.

    • Search Google Scholar
    • Export Citation
  • Doviak, R. J., S. Chen, and D. R. Christie, 1991: A thunderstorm-generated solitary wave observation compared with theory for nonlinear waves in a sheared atmosphere. J. Atmos. Sci., 48 , 87111.

    • Search Google Scholar
    • Export Citation
  • Droegemeier, K. K., and R. B. Wilhelmson, 1987: Numerical simulation of thunderstorm outflow dynamics. Part I: Outflow sensitivity experiments and turbulence dynamics. J. Atmos. Sci., 44 , 11801210.

    • Search Google Scholar
    • Export Citation
  • Fankhauser, J. C., N. A. Crook, J. Tuttle, L. J. Miller, and C. G. Wade, 1995: Initiation of deep convection along boundary layer convergence lines in a semitropical environment. Mon. Wea. Rev., 123 , 291313.

    • Search Google Scholar
    • Export Citation
  • Fovell, R. G., and P. S. Dailey, 2001: Numerical simulation of the interaction between the sea-breeze front and horizontal convective rolls. Part II: Alongshore ambient flow. Mon. Wea. Rev., 129 , 20572072.

    • Search Google Scholar
    • Export Citation
  • Frank, P. J., and P. A. Kucera, 2003: Radar characteristics of convection along colliding outflow boundaries observed during CRYSTAL-FACE. Preprints, 31st Int. Conf. on Radar Meteorology, Seattle, WA, Amer. Meteor. Soc., 12A.8.

    • Search Google Scholar
    • Export Citation
  • Fulton, R., D. S. Zrnic, and R. J. Doviak, 1990: Initiation of a solitary wave family in the demise of a nocturnal thunderstorm density current. J. Atmos. Sci., 47 , 319337.

    • Search Google Scholar
    • Export Citation
  • Intrieri, J. M., A. J. Bedard Jr., and R. M. Hardesty, 1990: Details of colliding thunderstorm outflows as observed by Doppler lidar. J. Atmos. Sci., 47 , 10811098.

    • Search Google Scholar
    • Export Citation
  • Karan, H., 2007: Thermodynamic and kinematic characteristics of low-level convergence zones observed by the mobile integrated profiling system. Ph.D. thesis. University of Alabama, 20 pp.

  • Karan, H., and K. R. Knupp, 2006: Mobile integrated profiler system (MIPS) observations of low-level convergent boundaries during IHOP. Mon. Wea. Rev., 134 , 92112.

    • Search Google Scholar
    • Export Citation
  • Kingsmill, D. E., 1995: Convection initiation associated with a sea-breeze front, a gust front, and their collision. Mon. Wea. Rev., 123 , 29132933.

    • Search Google Scholar
    • Export Citation
  • Kingsmill, D. E., and N. A. Crook, 2003: An observational study of atmospheric bore formation from colliding density currents. Mon. Wea. Rev., 131 , 29853002.

    • Search Google Scholar
    • Export Citation
  • Knupp, K. R., 2006: Observational analysis of a gust front to bore to solitary wave transition within an evolving nocturnal boundary layer. J. Atmos. Sci., 63 , 20162035.

    • Search Google Scholar
    • Export Citation
  • Koch, S. E., and W. L. Clark, 1999: A nonclassical cold front observed during COPS-91: Frontal structure and the process of severe storm initiation. J. Atmos. Sci., 56 , 28622890.

    • Search Google Scholar
    • Export Citation
  • Koch, S. E., P. B. Dorian, R. Ferrare, S. H. Melfi, W. C. Skillman, and D. Whiteman, 1991: Structure of an internal bore and dissipating gravity current as revealed by Raman lidar. Mon. Wea. Rev., 119 , 857887.

    • Search Google Scholar
    • Export Citation
  • Koch, S. E., C. Flamant, J. W. Wilson, B. M. Gentry, and B. D. Jamison, 2008: An atmospheric soliton observed with Doppler radar, differential absorption lidar, and a molecular Doppler lidar. J. Atmos. Oceanic Technol., 25 , 12671287.

    • Search Google Scholar
    • Export Citation
  • Mahapatra, P. R., R. J. Doviak, and D. S. Zrnić, 1991: Multisensor observation of an atmospheric undular bore. Bull. Amer. Meteor. Soc., 72 , 14681480.

    • Search Google Scholar
    • Export Citation
  • Mohr, C. G., L. J. Miller, R. L. Vaughn, and H. W. Frank, 1986: The merger of mesoscale data sets into a common Cartesian format for efficient and systemic analysis. J. Atmos. Oceanic Technol., 3 , 143161.

    • Search Google Scholar
    • Export Citation
  • Mueller, C. K., and R. E. Carbone, 1987: Dynamics of a thunderstorm outflow. J. Atmos. Sci., 44 , 18791898.

  • Oye, R., C. Mueller, and S. Smith, 1995: Software for radar translation, visualization, editing, and interpolation. Preprints, 27th Conf. on Radar Meteorology, Vail, CO, Amer. Meteor. Soc., 359–361.

    • Search Google Scholar
    • Export Citation
  • Ralph, F. M., C. Mazaudier, M. Crochet, and S. V. Venkateswaran, 1993: Doppler sodar and radar wind profiler observations of gravity-wave activity associated with a gravity current. Mon. Wea. Rev., 121 , 444463.

    • Search Google Scholar
    • Export Citation
  • Rottman, J. W., and J. E. Simpson, 1989: The formation of internal bores in the atmosphere: A laboratory model. Quart. J. Roy. Meteor. Soc., 115 , 941963.

    • Search Google Scholar
    • Export Citation
  • Seitter, K. L., 1986: A numerical study of atmospheric density current motion including the effects of condensation. J. Atmos. Sci., 43 , 30683076.

    • Search Google Scholar
    • Export Citation
  • Simpson, J. E., 1997: Gravity Currents in the Environment and the Laboratory. 2nd ed. Cambridge University Press, 244 pp.

  • Simpson, J. E., and R. E. Britter, 1980: A laboratory model of an atmospheric mesofront. Quart. J. Roy. Meteor. Soc., 106 , 485500.

  • Wakimoto, R. M., 1982: The life cycle of thunderstorm gust fronts as viewed with Doppler radar and rawinsonde data. Mon. Wea. Rev., 110 , 10601082.

    • Search Google Scholar
    • Export Citation
  • Wilson, J. W., and W. E. Schreiber, 1986: Initiation of convective storms at radar-observed boundary-layer convergence lines. Mon. Wea. Rev., 114 , 25162536.

    • Search Google Scholar
    • Export Citation
  • View in gallery
    Fig. 1.

    Upper air and surface observations on 16 Jun 2003. Solid and dashed contours depict isoheights/isobars and isobars/isotherms, respectively. The box shows the study area containing GBOS and KFSD WSR-88D radar. The line beginning at KFSD (K) and extending to the east represents 95 km distance where at the end convection initiation associated with synoptic forcing and BL circulations occurred. Soundings shown in Figs. 2, 7, and 21 were released from the MGLASS locations (G1 and G2) and the MIPS location (M).

  • View in gallery
    Fig. 2.

    Skew T–logp diagram acquired from MGLASS (G2) sounding at 2235 UTC. See Figs. 1 and 4b for location.

  • View in gallery
    Fig. 3.

    Time series of KFSD VAD wind profile on 16 Jun 2003. Full and half barbs represent 10 and 5 kt, respectively. The LCL height from MGLASS sounding (Fig. 2) is annotated.

  • View in gallery
    Fig. 4.

    (a) Sioux Falls, SD, WSR-88D radar reflectivity factor at 1 km AGL at 2235 UTC. The figure also depicts the eastward-moving boundary (B1), the slow westward-moving boundary (B2), and radar finelines indicative of boundary layer roll circulations, varying from 0 to 10 dBZ, (b) Geostationary Operational Environmental Satellite-12 (GOES-12) visible image taken at 2201 UTC. Locations of MGLASS sounding units (G1 and G2), KFSD radar site, and MIPS are annotated.

  • View in gallery
    Fig. 5.

    Radar finelines of boundary B1 in dashed (boundary 2 in solid) moving eastward (westward). Distance between each ring is 15 km. Boundary collision time is about 2345 UTC (1845 CST). The location of MGLASS-1 shown with a dashed line is about 110 km from KFSD.

  • View in gallery
    Fig. 6.

    Radar reflectivity factor (increasing light blue to red color with maximum values of 55–60 dBZe) overlaid on surface observations. The oval and parallelogram depict regions having different CI modes. The dashed line (also drawn in Fig. 1b) represents 95 km distance. The dark arrow indicates the first cell formation associated with a HCR, and the white arrow shows the first cell initiated by synoptic forcing. Instrument and boundary locations (B1, B2) are also annotated.

  • View in gallery
    Fig. 7.

    MGLASS 1 sounding at 2201 UTC on 16 Jun 2003 (see Fig. 4b for location G1). The sounding was released behind B1. The letter h indicates the height of the density current, and d marks the height of the coldest temperature; θυ−cold and θυ−env represents average virtual potential temperatures of cold outflow and environmental air over the depth of h = 740 m.

  • View in gallery
    Fig. 8.

    Surface observations with 1-min time resolution acquired at Mitchell, SD, ASOS (see Fig. 5 for the location); (upper) virtual temperature (Tυ, solid), mixing ratio (rυ, dotted), and pressure (p, dashed) variations. (lower) Wind speed (spd, dotted), and wind direction (dir, solid). The passage of B1 occurred during the 2137–2144 UTC time period.

  • View in gallery
    Fig. 9.

    Sioux Falls, SD, (KFSD) WSR-88D reflectivity factor on an xy plane at 0.5 km AGL at (a) 2330, (b) 2345 UTC 16 Jun, and (c) 0030 and (d) 0055 UTC 17 Jun 2003. Boundaries defined in the text are labeled. Here, S refers to an intense storm, and M represents the MIPS location.

  • View in gallery
    Fig. 10.

    Vertical structures of B1 and B2. The xz cross sections pass through the radar location (y = 0) as shown in Fig. 9; CB1 and CB2 are propagation speeds of B1 and B2, respectively. (a) Eastward-moving boundary B1. Arrows depict the u–w ground relative flow. Shaded contours represent B1-relative airflow normal to B1. (b) Westward-moving boundary, B2, and u–w ground relative flow (arrows). Negative shaded values (light gray) indicate an easterly boundary-relative flow. The solid contours in (a) and (b) are reflectivity factor in dBZ.

  • View in gallery
    Fig. 11.

    Ground-relative flow within the xz plane passing through y = 0 at (a) 2320, (b) 2330, and (c) 2335 UTC. Reflectivity factor contours are solid lines. Thicker reflectivity contours depict values greater than 12 dBZ. The KFSD radar is located at the origin. The vertical axis is stretched by a factor of two. Refer to Fig. 9 for relative location. The wind profiles on the right were obtained from G2 sounding (Fig. 2) and the KFSD VAD analysis at 2330 UTC between 0.6 and 3 km MSL.

  • View in gallery
    Fig. 12.

    Surface observations (School Net) with 1-min time resolution from Sioux Falls—Pavilion, SD, between 2200 UTC 16 Jun and 0200 UTC 17 Jun 2003. Vertical dashed lines indicate passage of boundaries B2, ACB1, and GF.

  • View in gallery
    Fig. 13.

    Same as Fig. 11, but during and shortly after the collision on 16 Jun 2003.

  • View in gallery
    Fig. 14.

    Same as Fig. 11, but at (a) 2355 16 Jun 2003, and (b) 0000 and (c) 0015 UTC 17 Jun 2003.

  • View in gallery
    Fig. 15.

    (a) Radial velocity at 0.5 km AGL. Dashed line depicts the cross section taken in Fig. 16 along the A–B (Fig. 9d), which has 20° angle from west. Negative values are toward the radar, outbound winds are positive. Arrows depict updraft–downdraft couplets associated with ACB2 and its secondary circulation; (b) Z reflectivity factor at 0.5 km AGL.

  • View in gallery
    Fig. 16.

    Same as Fig. 11, but for boundary ACB2 along the dashed line drawn in Figs. 9d and 15 at 0055 UTC 17 Jun 2003. Thick line depicts the kinematic boundary ACB2 associated with a solitary wave with wavelength λ of 7.2 km and height h1 of 1.4 km.

  • View in gallery
    Fig. 17.

    Radar reflectivity factor (Z) at 0.5 km AGL at (a) 0035 and (b) 0045 UTC 17 Jun 2003. The KFSD radar is located at x = 0, y = 0. The circle depicts the MIPS location.

  • View in gallery
    Fig. 18.

    MIPS surface observations between 0037 and 0105 UTC 17 Jun 2003. (a) Time series of virtual potential temperature, mixing ratio, and pressure (gray), (b) wind speed and wind direction (gray). The vertical dashed lines depict the timing of the updrafts measured by the 915-MHz wind profiler ahead of the GF. The arrow in (a) indicates in the relative dynamic pressure maxima, and in (b) the sounding launch time.

  • View in gallery
    Fig. 19.

    Time–height section of 915-MHz profiler vertical beam measurements (a) SNR and (b) w acquired at 1-min intervals. (b) Values in black and white colors represent downdrafts exceeding 4 m s−1. (a),(b) Ceilometer-derived cloud-base height (black) is shown as solid dots. (c) Surface pressure time series. (d) Time–height section of ceilometer two-way attenuated backscatter profile. Values exceeding the color bar are in black. The inset in upper-right corner displays backscatter values between 0043 and 0053 UTC from surface to 0.7 km. (e) Time–height observations of water vapor mixing ratio acquired from the Microwave Profiling Radar (MPR). Measurements prior to 0051 UTC are not accurate because of warm up time. Mixing ratio contours are drawn for every 1 g kg−1 starting at 6 g kg−1. (f) Time series of mean vertical velocity acquired from 915-MHz profiler at 1 (solid) and 2.6 km (dotted) AGL. The CBZ passage occurred between 0043 and 0049 UTC. The solid vertical line represents the time of the sounding (0053 UTC) shown in Fig. 21.

  • View in gallery
    Fig. 20.

    (a) Time–height section of horizontal winds acquired by the 915-MHz profiler between 0030 and 0130 UTC. The solid gray line outlines the approximate gust frontal head region, and the dashed line defines the approximate top of the trailing density current. (b) The 2-min average winds derived from the Doppler sodar. The solid vertical line represents the time of the sounding (0053 UTC) shown in Fig. 21. Half and full barbs represent 2.5 and 5 m s−1 winds, respectively.

  • View in gallery
    Fig. 21.

    Skew T–logp diagram acquired from the sounding released by the MIPS crew at 0053 UTC after the gust frontal passage. Mixing ratio lines are depicted in dotted lines slanted from left to right; θυ–cold is the average virtual potential temperature over the depth of 400 m; Tυ = 29°C is the virtual temperature prior to the GF passage.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 189 107 3
PDF Downloads 86 21 0

Radar and Profiler Analysis of Colliding Boundaries: A Case Study

Haldun KaranUniversity of Alabama in Huntsville, Huntsville, Alabama

Search for other papers by Haldun Karan in
Current site
Google Scholar
PubMed
Close
and
Kevin KnuppUniversity of Alabama in Huntsville, Huntsville, Alabama

Search for other papers by Kevin Knupp in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

The kinematics of a head-on collision between two gust fronts, followed by a secondary collision between a third gust front and a bore generated by the initial collision, are described using analyses of Weather Surveillance Radar-1988 Doppler (WSR-88D) and Mobile Integrated Profiling System (MIPS) data. Each gust front involved in the initial collision exhibited a nearly north–south orientation and an east–west movement. The eastward-moving boundary was 2°C colder and moved 7 m s−1 faster than the westward-moving boundary. Two-dimensional wind retrievals reveal contrasting flows within each gravity current. One exhibited a typical gravity current flow structure, while the other assumed the form of a gravity wave/current hybrid with multiple vortices atop the outflow. One of the after-collision boundaries exhibited multiple radar finelines resembling a solitary wave shortly after the collision. About 1 h after the initial collision, a vigorous gust front intersected the eastward-moving bore several minutes before both circulations were sampled by the MIPS. The MIPS measurements indicate that the gust front displaced the bore upward into a neutral residual layer. The bore apparently propagated upward even farther to the next stable layer between 2 and 3 km AGL. MIPS measurements show that the elevated turbulent bore consisted of an initial vigorous wave, with updraft/downdraft magnitudes of 3 and −6 m s−1, respectively, followed by several (elevated) waves of decreasing amplitude.

* Current affiliation: Stennis Space Center, Stennis Space Center, Mississippi.

Corresponding author address: Haldun Karan, Stennis Space Center, Bldg 1103, Rm. 233, Stennis Space Center, MS 39529. Email: karan@ngi.msstate.edu

Abstract

The kinematics of a head-on collision between two gust fronts, followed by a secondary collision between a third gust front and a bore generated by the initial collision, are described using analyses of Weather Surveillance Radar-1988 Doppler (WSR-88D) and Mobile Integrated Profiling System (MIPS) data. Each gust front involved in the initial collision exhibited a nearly north–south orientation and an east–west movement. The eastward-moving boundary was 2°C colder and moved 7 m s−1 faster than the westward-moving boundary. Two-dimensional wind retrievals reveal contrasting flows within each gravity current. One exhibited a typical gravity current flow structure, while the other assumed the form of a gravity wave/current hybrid with multiple vortices atop the outflow. One of the after-collision boundaries exhibited multiple radar finelines resembling a solitary wave shortly after the collision. About 1 h after the initial collision, a vigorous gust front intersected the eastward-moving bore several minutes before both circulations were sampled by the MIPS. The MIPS measurements indicate that the gust front displaced the bore upward into a neutral residual layer. The bore apparently propagated upward even farther to the next stable layer between 2 and 3 km AGL. MIPS measurements show that the elevated turbulent bore consisted of an initial vigorous wave, with updraft/downdraft magnitudes of 3 and −6 m s−1, respectively, followed by several (elevated) waves of decreasing amplitude.

* Current affiliation: Stennis Space Center, Stennis Space Center, Mississippi.

Corresponding author address: Haldun Karan, Stennis Space Center, Bldg 1103, Rm. 233, Stennis Space Center, MS 39529. Email: karan@ngi.msstate.edu

1. Introduction

This study is concerned with the collision of near-surface convergent boundary zones (CBZs), which are characterized by a pressure increase, a change in wind direction or wind speed, low-level convergence, and updrafts aloft. A variety of CBZs within the atmospheric boundary layer (ABL) have been observed in previous studies. These include 1) thunderstorm outflow boundaries, or gust fronts (GFs), 2) classic mesoscale circulations driven by horizontal gradients in density such as sea breeze fronts, 3) large eddies in the convective boundary layer (CBL) such as horizontal convective rolls (HCRs), 4) synoptic-scale warm and cold fronts (deep and shallow), prefrontal troughs, and drylines, 5) gravity waves, 6) bores and solitary waves, and 7) leeside convergence zones. The above examples possess a broad range of spatial/temporal scales, depths, vertical motion characteristics, and associated cloud features. In this article, we consider the kinematics of a collision between two gust fronts moving in opposite directions, followed by the collision of a third new gust front with a bore produced by the initial collision.

Internal gravity waves and atmospheric undular bores have been studied with the aid of numerical models and theories (Clarke 1984; Doviak et al. 1991; Crook and Miller 1985; Crook 1986, 1988), and observations (Clarke et al. 1981; Doviak and Ge 1984; Cheung and Little 1990; Fulton et al. 1990; Doviak et al. 1991; Mahapatra et al. 1991; Ralph et al. 1993; Knupp 2006; Coleman et al. 2009). Koch and Clark (1999) investigated a bore formation along a cold front and its propagation ahead of the cold front on a low-level nocturnal inversion. In this and the other cited studies, the propagation of a density current into a stratified-stable layer capped with temperature inversion, or the collision of two density currents, is often conducive to formation of internal gravity waves/bores. Boundary collisions have drawn attention because of their apparent high efficiency in convective initiation (CI), cloud/storm mergers, and atmospheric bore generation (Kingsmill and Crook 2003). Using laboratory tank experiments, Simpson (1997) examined the collision of two gravity currents of different size and density. At the time of the collision, the two gravity currents were reflected upward. Shortly after the collision, two bores traveling in opposite directions formed within the stable air of each outflow. Kingsmill and Crook (2003) observed similar structures in the form of radar finelines in seven out of 10 cases involving collisions between outflows and the sea breeze front over Florida. Their study suggested a clear relationship between virtual potential temperature deficits within the two density currents and the after-collision characteristics.

Lidar observations of small-scale interactions of colliding outflows (Intrieri et al. 1990) revealed that the warmer of the two outflows was deflected upward to heights of 2 km by the colder outflow. This mechanical lifting produced CI in all cases examined. Wilson and Schreiber (1986) observed that colliding convergence lines initiated new storms or intensified existing storms in 71% of the cases in Colorado. Yet, Kingsmill (1995), after examining gust front–sea breeze collisions, did not observe an enhancement in convection associated with boundary collisions. The outcome of boundary collisions appears to be dependent upon several parameters, including collision angle, boundary propagation direction, depth of the colliding boundaries, and vertical profile of virtual potential temperature deficit (density) within each density current. Frank and Kucera (2003) investigated several parameters for their role in convective initiation associated with collision of gust fronts. They found that in six out of eight cases, convective initiation occurred when the angle between the two boundaries was less than 40°. They also reported that nine out of 14 collision cases initiated convection when boundaries moved in opposite directions.

The current study investigates an event in which two outflow boundaries collided very close to the University of Alabama in Huntsville (UAH) Mobile Integrated Profiling System (MIPS) and the Weather Surveillance Radar-1988 Doppler (WSR-88D) radar in Sioux Falls, South Dakota (KFSD). The study area and the synoptic settings are presented in sections 2 and 3. Kinematic and thermodynamic characteristics of the ABL and the boundaries are discussed in the sections 4 and 5. Gust-front collision and postcollision characteristics are discussed in sections 6. The MIPS observations of a gust front passage and its interaction with one of the after-collision boundaries are presented in the section 7. Conclusions follow in section 8.

2. Study area and the data

The Bow Echo and Mesoscale Convective Vortices Experiment (BAMEX) field study was designed to examine bow echoes and mesoscale convective vortices (Davis et al. 2004). The BAMEX mobile ground-based observational platforms acquired measurements on 16 June 2003 in advance of a weakening mesoscale convective system (MCS) over southeastern South Dakota. These included the MIPS, the Mobile Meteorological Measurement Vehicle (M3V), and two National Center for Atmospheric Research (NCAR) GPS loran atmospheric sounding system (MGLASS) units. The topography over the analysis region (box in Fig. 1b) ranges from 400–550 m above mean sea level (MSL).

The MIPS instruments utilized in this study include a 915-MHz radar wind profiler, a 2-KHz Doppler sodar, a 12-channel microwave profiling radiometer, a 905-nm ceilometer, and surface instrumentation. Further details on the MIPS instrument characteristics are provided in Karan and Knupp (2006) and online at http://vortex.nsstc.uah.edu/mips.

Additional high-resolution (1 min) surface data were available from Automated Surface Observing System (ASOS) and school network surface stations around Sioux Falls. The WSR-88D radar was operating in VCP11 volume scan mode (0.5°, 1.5°, 2.4°, 3.4°, 4.3°, 5.3°, 6.2°, 7.5°, 8.7°, 10°, 12°, 14°, 16.7°, and 19.5° elevation angles). Radar reflectivity factor (Z) and radial velocity (Vr) data from the Sioux Falls WSR-88D radar are used to infer locations of boundaries, their propagation speeds, and their time-dependent kinematic structures. Following editing with the SOLO-II software, Vr and Z were interpolated to a Cartesian grid using the NCAR REORDER software (Oye et al. 1995). Grid spacing was 0.25 km in the x, y, and z directions. The NCAR Custom Editing and Display of Reduced Information in Cartesian space (CEDRIC; Mohr et al. 1986) software package was used to synthesize vertical and horizontal velocities and Z field before, during, and after the collision of the boundaries.

Unlike the kinematic structures around drylines, density currents—more specifically, shallow and denser thunderstorm outflows—have been shown and observed to align themselves perpendicular to their attended fast-moving boundaries. Wakimoto (1982) analyzed time-dependent thunderstorm gust-front structures with the use of single Doppler radars. The author stated “the winds behind the gust fronts are generally perpendicular to the frontal interface, so that a radar pointing normal to the front would record Doppler velocities that are very close to the actual horizontal velocities at low elevation angles.” Numerous observations have observed gust frontal outflows that were normal to the gust front (Charba 1974; Mueller and Carbone 1987; Karan and Knupp 2006). Dual-Doppler wind analyses (e.g., Mueller and Carbone 1987) show a sharp quasi–two dimensional convergence zone with winds relatively normal to the gust-front boundaries at low levels (<1.2 km). Using lidar observations, Intrieri et al. (1990) also made the same 2D assumption to infer the kinematic characteristics associated with boundaries and their collisions. They performed upward integration of the anelastic continuity equation to derive vertical velocities. Knupp (2006) also assumed 2D mass continuity to infer w within a CBZ that exhibited a gust-front–bore–solitary wave transition. Therefore, flow structures around the vicinity of the two outflow boundaries are deduced by assuming that the flow is two dimensional within the vertical plane normal to the boundary.

In the present case, the validity of the 2D assumption was tested by examining the Vr fields along the CBZ during their movement over the radar. In general, variations in Vr are found to be one order of magnitude less than variations in Vr normal to the boundary, thus indicating that the 2D assumption is accurate to with about 10%. Even for slightly nonnormal angles between the radar radial and boundary, the 2D analysis is still robust, provided that the wind direction does not vary substantially across the boundary, as was the case here. Using two-dimensional mass continuity equation, error calculations in vertical velocities at 1 km AGL prior to the collision time varied from 4% to 16% within ±5 km along (north–south direction intersecting the x axis; e.g., Fig. 9) and 15 km on either side of the boundaries.

3. Large-scale setting

Upper-level flow at 1200 UTC over the experimental domain (shown as a square in Fig. 1) was weak over much of the troposphere. A ridge at 500 hPa (Fig. 1a) was located over the western United States. Patterns at 850 hPa (not shown) reveal weak advection of water vapor and warm air from the south into eastern South Dakota.

Figure 1b shows 1800 UTC surface observations and locations of the KFSD WSR-88D radar (K), MIPS (M), and NCAR MGLASS mobile sounding units (G1 and G2, all within the square). Weak southeasterly flow over western Iowa and eastern Nebraska was located southeast of a weak low pressure center over northern South Dakota. A region of large-scale confluence is apparent in the surface wind field over and east of the study area. The 6-h Weather Research and Forecast model (WRF-12) forecast initialized at 1200 UTC indicated convective available potential energy (CAPE) values of 1000–1500 J kg−1 and convective inhibition (CIN) of 50–75 J kg−1 over the region. An MGLASS sounding launched from point G2 at 2235 UTC (Fig. 2) indicates a well-mixed CBL extending to about 2 km above ground level (AGL). Throughout the depth of the CBL, the flow was southwesterly at 5–6 m s−1. Weak northerly flow existed above the CBL between 650 and 450 hPa. Mid- and upper-tropospheric winds (northerly and westerly, respectively) were less than 3 and 6 m s−1, respectively, indicative of weak synoptic forcing. Surface based CAPE and CIN values of this sounding were only 420 and 10 J kg−1, respectively.

4. ABL properties prior to collision

A time series of velocity azimuth display (VAD) winds (Fig. 3) provides a good portrayal of the mean mesoscale wind profile over the analysis domain within and above the CBL. The wind exhibits strong veering from 160° at 0.6 km MSL (0.15 km AGL) to about 230° at 2.6 km (2.15 km AGL), the estimated CBL top (Fig. 2). The mean flow within the CBL was from the southwest at about 6 m s−1. Appreciable directional wind shear is indicated within weak flow above the CBL.

The locations of the two outflow boundaries (B1, B2) and horizontal convective roll circulations are represented by radar fine lines as annotated in Fig. 4a. Cumulus cloud streets were produced by HCR updrafts (Fig. 4b). The HCR orientation was approximately aligned with the low-level wind and shear direction. Boundary B1 systematically encountered HCR updrafts and downdrafts as it propagated to the east. Although intersections of HCRs with boundaries have been shown to be preferred locations for convective initiation (Fankhauser et al. 1995; Atkins et al. 1998; Fovell and Dailey 2001), the KFSD WSR-88D data indicate that interactions between B1 and HCR did not initiate deep convection, although relatively deep nonprecipitating cumulus clouds with tops to about 4.5 km AGL can be inferred from the Doppler analyses presented in section 5a (Fig. 11c).

5. Evolution of B1 and B2

A time series of isochrones of B1 and B2 from 2200 to 2345 UTC is shown in Fig. 5. Locations of key surface stations, the KFSD radar, the MGLASS sounding units, and the MIPS are also annotated. The radar fine line associated with the eastward-moving boundary (B1, dashed line), which was produced by a larger MCS, first appeared 100 km west of the KFSD radar around 2145 UTC. It moved eastward at 11.5 m s−1 between 2200 and 2345 UTC.1 In contrast, the westward-moving boundary (B2) was an older, gust front. It formed 60 km east of the KFSD radar at 1845 UTC from a merger of several outflow pools. Its average ground-relative speed over the same time period was only 4.2 m s−1. Just prior to collision at 2345 UTC, the propagation speed of B2 (4.2 m s−1) remained unchanged, while the 10-min average speed of B1 had decreased to 8.5 m s−1.

The first convective initiation within the study area commenced at 1730 UTC, 65 km east of the KFSD radar along an HCR axis. This cell is identified with the black arrow in Fig. 6a. Surface observations (Fig. 1) indicate large-scale confluence within a north–south region, labeled as parallelogram A. This cell dissipated within 30 min, but two gust fronts remained. One moved eastward and the other (boundary B2) moved westward as shown in Fig. 6b. Boundary B2 moved against a component of the ambient wind (and vertical wind shear) within the boundary layer, and initiated numerous [O(50)] convective cells between 2045 and 2345 UTC. In contrast, B1 initiated only one short line of deep convection (southeast of label “G1” in Fig. 6d) as it moved eastward between 2200 and 2345 UTC.

a. Evolution and structure of B1

1) Thermodynamics

A sounding from the MGLASS-1 (see Figs. 4b and 6 for the location) was acquired at 2201 UTC within the trailing density current of B1, 105 min prior to collision. Figure 7 shows profiles within the cold air over the 740 m depth of the outflow, indicated by the shaded region (wind was not available). The coldest air was confined to the lowest 300 m. Using this sounding and the observed ground-relative propagation speed (Vgf = 11 m s−1), the Froude number (Fr) was estimated using the relations
i1520-0493-137-7-2203-e1
where uamb is the ambient flow along the direction of gust-front propagation over the depth (h = 740 m) of the density current, b is an experimentally determined adjustment factor (0.62, included to account for the airflow in advance of the density current) determined from the laboratory experiments of Simpson and Britter (1980), and θυc (301.6 K) and θυw (305.1 K) are the respective virtual potential temperatures averaged over the depth h within the density current and in advance of the density current. Insertion of these measured parameters into Eq. (1) yields Fr = 1.3, close to the theoretical upper limit of 2 (Benjamin 1968).
The ASOS surface station at Mitchell, South Dakota, also recorded the gust-front passage 24 min earlier at 2137 UTC (Fig. 8). The virtual temperature dropped 3.5°C (32.5° to 29°C) and the mixing ratio increased from 11.4 to 12.8 g kg−1. Wakimoto (1982) has shown that maximum surface pressure rise is within 10% of the value calculated hydrostatically. Therefore, the observed surface pressure increase can be used to estimate the Froude number of the density current according to the relation (Seitter 1986)
i1520-0493-137-7-2203-e2
where Δp is the assumed hydrostatic pressure increase (∼0.5 hPa) measured at surface. Surface density prior to the boundary arrival is used for ρw, even though it is defined as an average value of environmental atmospheric density over the depth of cold outflow. The observed ground-relative propagation speed and these surface observations applied to Eq. (2) produce a Froude number of 1.4, in close agreement with the 1.3 value computed from Eq. (1).

2) Kinematics

Since both B1 and B2 exhibited an approximately north–south orientation and east–west directional propagation close to the KFSD radar (Figs. 6 and 9a), a two-dimensional Doppler analysis was performed using the east–west component of the flow (u component) along the y = 0 axis (dashed line in Fig. 9a), assuming that radial velocities along 270° and 90° azimuth represent the u component of the flow and that the boundary did not exhibit significant variability along its major axis (y axis in this case).2 Vertical air motion (w) was calculated from the two-dimensional anelastic continuity equation, ∂(ρu) + ∂(ρw) = 0 (ρ is air density), applying a w = 0 boundary condition (variational constraint) at both the surface and the level at which reflectivity factor was less than −5 to −10 dBZ.

The kinematic structure of boundary B1 5 min before the collision is shown in Fig. 10a. At this time, B1 exhibited a deep updraft about 2 km wide (excluding the weaker ascent diagnosed for x > −17 km), with a maximum of 7–8 m s−1 located along the forward side of the reflectivity gradient (near x = −19). The positive feeder flow of 2 m s−1 behind B1 (u′, relative to the movement speed of 10 m s−1, and indicated by the gray shading) is indicative of mass transport (below 0.6 km in this vertical section), which is characteristic of density currents. The MGLASS-1 sounding discussed in the previous section observed this gust-front passage with 740-m-deep outflow about 1.5 h earlier.

The kinematic analysis in Fig. 11 portrays an interaction between B1 and an HCR marked by enhanced Z near x = −21. At 2320 UTC, B1 exhibited respective maxima in Z and updraft of 15 dBZ and 7 m s−1 near x = −29.5 km and a height of 0.8 km AGL (Fig. 11a). Updrafts within B1 at 2320 and 2330 UTC tilted eastward (ground-relative frame), in response to the westerly flow component within the CBL (right-hand panel of Fig. 11). Interaction with the HCR at 2330 (Fig. 11b) resulted in two adjacent reflectivity maxima near x = −21 and x = −25 km. The HCR and B1 updrafts merged by 2335 into a single intense (7–8 m s−1) updraft and corresponding single 18 dBZ reflectivity core (Fig. 11c). Since this updraft extends to about 4.5 km AGL, which is about 2.5 km above the lifting condensation level (LCL) of surface air (Fig. 2), this updraft represents a nonprecipitating cumulus congestus cloud. An easterly return flow west of the updraft and above 3 km is apparent west of the updraft.

b. Evolution and structure of B2

The passage of B2 over the Sioux Falls—Pavilion surface station (5.5 km south of KFSD radar; see Fig. 5 for location), produced a virtual temperature deficit of only 1.6°C (1.4°C less than that sampled in B1) and an increase in rυ of 1.4 g kg−1 around 2245 UTC (Fig. 12). The surface virtual temperature behind B2 was 30.8°C, 1.8°C warmer than that within the B1 outflow. Unfortunately the pressure sensor at this School Net site was not sensitive enough to resolve the relatively small surface pressure change. An average propagation speed of B2 during the 2200–2345 UTC time period was 4.2 m s−1. The sounding G2 (Fig. 2) was used to determine the average ambient wind and virtual temperatures from surface to 500-m height. The average virtual temperature was assumed to be linearly decreasing within the lowest 500 m. The depth of B2 (∼500 m; Fig. 10b) was applied to Eq. (1), which yields Fr of 1.6, above the upper limit of the theoretical Fr for density currents.

The kinematic structure of boundary B2 is best illustrated (relative to the structure of B1) in Fig. 10b. In contrast to B1, B2 did not exhibit significant mass transport (a system relative feeder flow) behind its head (u′ ≈ 0 m s−1). Although B2 exhibited higher reflectivity values (∼23 dBZ) the maximum updraft of 5 m s−1 was 1–2 m s−1 less than that within B1. The updraft width (3 km) appears to be greater than that diagnosed in B1. That the radar fineline in B2 is both wider and more intense than in B1 suggests that B2 was more effective in concentrating and lofting insects upward into the boundary layer. This is consistent with the greater number of CI events produced by B2. Above 1 km AGL, the relative flow overturns toward the east behind the gust frontal head, creating a return flow that descends and feeds the outflow of B2 in the form of a rotor circulation (Fig. 10b). This rotor was a persistent feature of B2. Mueller and Carbone (1987) found that significant vertical shear along the interface of two fluids is conducive to shear instability, the creation of inflection points, and a subsequent evolution to vortex rolls.

6. Gust-front collision and postcollision characteristics

a. Radar analysis of the gust-front collision

Figure 9 illustrates a plan view of the collision process from 15 min before to 70 min after collision. Maximum updraft strengths within B1 and B2 increased from 5–6 m s−1 to 7–8 m s−1 between 2330 and 2345 UTC. During and after the collision, new cells were initiated at isolated points along the collision line 30–40 km south and north of the collision boundary axis. A boundary identified as ACB2 (after-collision boundary of B2) in Figs. 9c,d appears as two parallel radar finelines (along the dashed line AB in Fig. 9d) exhibiting wavelike characteristics. Intense deep convection located 30–40 km southwest of the KFSD radar (denoted with a letter S) formed a new gust front that propagated toward the northwest to east. Figure 9d reveals that the east-moving segment of this gust front intersected boundary ACB1 (after-collision boundary B1). Details of the structure of this gust front and the modified ACB1 are examined with MIPS measurements in section 7.

The kinematic structures of B1 and B2 about 5 min prior to collision are presented in Fig. 13a. The main characteristics of B1 include a deep updraft, and an oscillatory return flow at 3–4 km AGL west of the updraft. In contrast, B2 exhibits two horizontal vortices east of its main updraft by 2345 UTC (Fig. 13b). Thus B2 resembles a gravity current and gravity wave combination as opposed to the pure gravity current structure of B1.3 As shown earlier by means of Froude number calculations [Eq. (1)], B2 was not a pure density current, but rather was a hybrid phenomenon of gravity current-gravity wave.

By 2345 UTC the updrafts of B1 and B2 had completely merged near x = −12.5 km (Fig. 13b) into a stronger and wider updraft and reflectivity core. At the time of collision, the convergent zone was wider and stronger (between −4 × 10−3 and −8 × 10−3 s−1), and the depth of convergence increased to 1.2 km AGL. The updraft exhibits a maximum value of 12.5 m s−1 at 1.2 km AGL and extends to 4 km AGL. (Clouds were observed along this convergence line.) Reflectivity values greater than 12 dBZ also reached maximum heights of 1.6 km AGL at this time. These observations confirm significant enhancement in updraft intensity and horizontal width (during the collision event) as observed by Kingsmill and Crook (2003) and simulated in the laboratory by Simpson (1997).

At 2350 UTC, 5 min after the collision, two separate updrafts emerged on each side of the Z maximum, one centered at x ≈ −10.5 km with a wmax of 7 m s−1 and the other at x ≈ −13 km with a wmax of 4 m s−1 (Fig. 13c). Even though the density was not identical within each gravity current, we note that two surface rooted CBZ, propagating in opposite directions, resulted from the collision event.

b. Radar analysis of the postcollision flows

Two separate updrafts with individual Z maxima are apparent 10 min after the collision at 2355 UTC in Fig. 14a. Like its predecessor B2, ACB2 displayed a steep updraft with a wmax of 6 m s−1 near x = −13 km. Although the updrafts associated with ACB1 (x = −9 km) exhibit a tilt similar to that of the B1 updrafts, the subsequent evolution is ambiguous due to lack of complete radar coverage at close radar range. The two postcollision boundaries are separated with a broad downdraft peaking at −2 m s−1. It is noteworthy that the updrafts of ACB1 are surface rooted to low-level convergence shortly after the collision—at least in this vertical plane. The analysis of MIPS data presented in section 7 reveals a contrasting structure—an elevated wave confined to the 2–3 km AGL layer—about 1 h later and 40 km south of the analysis presented in Fig. 14a.

During the period between 2355 and 0015 UTC, ACB1 propagated faster than ACB2 (Fig. 14). The reflectivity fields associated with each boundary decreased in height and increased in width. The maximum w values within ACB2 and ACB1 were about 5 and 3.5 m s−1, respectively. The outflow of B1 created a stable layer with uniform flow indicated west x = −20 km in Fig. 14c. As ACB2 propagated into the residual outflow of B1, two successive updraft–downdraft couplets formed behind ACB2. Radial velocity observations (not shown) suggest that an outbound flow associated with ACB2 was first detected at 2355 UTC between 0.4 and 0.8 km AGL. The flow within ACB2 at 0000 UTC resembles a hydraulic jump that characterizes a bore (e.g., Knupp 2006); that is, air parcels rise in the updraft and remain elevated, exhibiting wave oscillations within the westerly flow between 1 and 2 km AGL. However, a more prominent up–down flow 15 min later at 0015 UTC (Fig. 14c) presents a pattern more consistent with that of a solitary wave, with a primary crest centered near x = −15 km and a secondary crest at x = −8 km. Both the updrafts and downdrafts within the westernmost wave exhibit a magnitude of about 5 m s−1. The depression of the Z contours within reduced Z near x = −13 is physically consistent with this interpretation. This relative minimum clearly separates maxima in Z associated with the crest of each wave.

The secondary updraft associated with the ill-defined vortex later intensified and developed a more prominent circulation, completely separated from the circulation of ACB2. ACB2 propagated westward at a faster rate, 5–6 m s−1, than its earlier form (B2). It moved 60 km west-northwest of KFSD before disappearing from radar around 0200 UTC. Unfortunately, there were no surface stations along the path of the ACB2 to characterize its thermodynamic features.

Figure 15 shows that two lines of positive radial velocity were apparent by 0055 UTC. Their separation distance and average westward speed are estimated at about 8.2 km (similar to the spacing at 2355 UCT) and 5.5 m s−1, respectively. The north-pointing arrows in Fig. 15 depict locations of updrafts associated with convergent boundaries for each wave. A vertical section through these features (Fig. 16) reveals maximum low-level westerly flow (7 m s−1) west of ACB2, and easterly flow (−4 m s−1) that trails this feature (east of x = −20). Therefore, ACB2 resembles a dual-crested solitary wave consisting of a downdraft (2–3 m s−1) separating two updrafts of about 5 m s−1 intensity.

This scenario reveals a greater complexity than that indicated in idealized laboratory simulations (Simpson 1997) and observations (Kingsmill and Crook 2003). The height of the solitary wave (h1) and the wavelength (λ) seen in Fig. 16 were observed to be about 1.4 and 7.2 km, respectively. The type of the bore structure depends on the Froude number, Fr, and the ratio d0/h0, where d0 is the depth of the cold outflow and h0 is the prefrontal height of the stable layer (Rottman and Simpson 1989). As seen earlier, the depth of the most stably stratified layer, prior to the B1 passage, was 0.75 km (h0). Figure 10b suggests that the depth of B2 is 0.5 km. One of the difficulties specific to this case was to determine the height of the collided boundary seen at Fig. 13b. Upward displacements of enhanced reflectivity values suggest that the mixed boundary at the time of the collision was lifted to the height of 1.5 km AGL at 2345 UTC. There was a considerable turbulent mixing between the two boundaries at that time. Therefore, the depth of the boundary ACB2 during the collision period was roughly between 1–1.2 km AGL. The ratio, d0/h0, then becomes 1.3–1.6. To compute Fr as B2 passed over Sioux Fall surface station (Fig. 12), reduced gravity, g′, was computed by using averaged virtual potential temperatures within the lowest 0.5 km layer based upon the sounding MGLASS-G2 at 2235 UTC for the ambient air and surface observations and adiabatic lapse rate assumption behind the boundary B2. Considering boundary normal westerly wind of 2.2 m s−1, Eq. (1) results in Fr number of 0.9. Simpson (1997) describes the bore character with respect to the ratio between the bore’s mean depth and depth of the stable layer. After these values were plotted on the parameter space between Fr and d0/h0, the bore strength h1/h0 are found to be 1.9–2.0, which according to laboratory bore simulations (Simpson 1997) indicates type A, smooth undular, bore (see also Figs. 3a,b of Rottman and Simpson 1989). Then, the bore depth can be deduced from the following relation: db/h0 = the bore strength ∼1.9–2.0. The theoretical values of the bore height yield values of 1.4–1.5 km, which is in close agreement with the observed bore height seen in Fig. 16. The radar reflectivity field confirmed that this dual radar finelines disappeared within the two hours from their formation. A close agreement between observed mean bore depth and the bore depth from hydraulic theory has been documented by previous observational and numerical studies (Koch et al. 1991; Ralph et al. 1993; Koch and Clark 1999; Knupp 2006; Koch et al. 2008).

7. MIPS observations

While ACB1 continued to move eastward, a new outflow was produced by a strong downdraft and associated 25 m s−1 horizontal wind gust (measured by the M3V just outside the reflectivity core) associated with storm “S” at about 0025 UTC. A radar fineline moved away from the storm in an arc from the east to the northwest (Fig. 17). The MIPS sampled this gust front at 0046 UTC and the passage of a modified form of ACB1 at 0050 UTC. Figure 17 indicates that ACB1 was superimposed on the GF boundary north of the MIPS by 0045 UTC. Further south, just west of the MIPS, ACB1 was diffuse and relatively weak.

a. Surface and general fineline characteristics

MIPS surface observations (Fig. 18) indicate that pressure steadily increased in the presence of weak southerly flow, while virtual potential temperature and mixing ratio remained nearly constant at 305 K and 10.5 g kg−1, respectively, during the 0035–0042 UTC period. A 90° wind shift to 270°, and wind gusts to 14 m s−1, (∼3 m s−1 greater than boundary propagation speed) accompanied the GF passage at 0046 UTC. Based on surface observations, the CBZ passage occurred between 0041 and 0049 UTC. This time interval is also characterized by the updraft and downdraft branches of the GF measured by 915 MHz profiler. Pressure and virtual potential temperature across the CBZ varied by +0.64 hPa and −1.4 K, respectively. The initial pressure increase of 0.36 hPa is attributed to dynamic effects,4 while the subsequent increase of 0.28 hPa following the GF passage represents the hydrostatic component (Charba 1974; Wakimoto 1982; Droegemeier and Wilhelmson 1987).

b. Vertical structures of the boundaries

MIPS 915-MHz wind profiler observations between 0037 and 0133 UTC are shown in Fig. 19. The profiler sampled a low-level updraft–downdraft circulation associated with the GF, followed by an intriguing w oscillation near 2.5 km AGL, apparently associated with ACB1. After the passage of the GF, a prolonged enhancement in 915-MHz signal-to-noise ratio (SNR) (Fig. 19a) and in ceilometer backscatter (Fig. 19d) was measured below 500 m AGL, and perturbations in SNR are associated with the w field below 1.5 km. Horizontal flow derived the 915-MHz profiler (Fig. 20) shows westerly flow within the GF head, followed by a much shallower layer of westerly flow below 0.5 km AGL that coincides with the enhanced 915-MHz SNR. Relatively strong southwest (SW) flow was present above the (distinct) top of the density current. Thus, this pattern is consistent with the shallow density current indicated by the 915 and ceilometer backscatter (Figs. 19a,d).

The w pattern within the gust front (Fig. 19b) is unusual in two respects: (i) The updraft is not contiguous and first appears aloft within the 0.5–1.5 km layer, and then below 0.5 km after 0045 UTC, just before the GF passage at 0046 UTC. (ii) The trailing downdraft centered near 1 km AGL has a magnitude of 6 m s−1 that is considerably stronger than that of the updraft (2.6 m s−1), and is immediately adjacent to the updraft. This latter feature is in contrast to a more typical pattern (based on a comprehensive analysis of gust fronts summarized by Karan 2007) in which the trailing gust front downdraft follows a narrow, vertically continuous updraft by 2–10 min. While it is tempting to interpret the oscillation in w at the 1 km level as an internal gravity wave (see Fig. 19f for greater details in the w time series at 1.0 km AGL), a sounding launched 7 min later shows a nearly neutral stratification at this level. However, the lowest several hundred meters was rendered stable by the earlier passage of B2 and cloud shading [e.g., the surface temperature prior to the gravity wave (GW) passage was about 27°C, 4°C less than the maximum temperature; see Fig. 21] so this stable layer would have been lifted by the gust front. Moreover, the time series of surface pressure (Fig. 19c) is in phase with that expected from the w pattern at 1 km (e.g., relative high and low pressure beneath respective ascending and descending portions of the wave).

The characteristics of ACB1, as indicated in Fig. 19, differ substantially from that of ACB2 and the earlier form of ACB1, both of which assumed a surface-based bore or solitary wave structure (see Fig. 14). Figure 19a indicates that significant variations in 915 SNR and w above 1.6 km (the top of the residual layer) were associated with the passage of ACB1 over the MIPS around 0050 UTC. A relative maximum in SNR within the 2–3 km AGL layer between 0050 and 0055 UTC is collocated with a wavelike oscillation around ACB1. Although the initial w oscillation is most significant (updraft of 2.5 m s−1, downdraft of −5 m s−1), subsequent oscillations in both SNR and w are evident through 0125 UTC within this layer following the passage of ACB1 (see Fig. 19f for more details of the w time series at 2.6 km AGL). The arrival of ACB1 was associated with a sudden 0.5 km increase in cloud-base height. The significant variations in cloud-base height (indicated in Fig. 19d and also drawn as black dots in Figs. 19a,b) suggest that the w oscillations within this layer were quite turbulent in character.

A balloon sounding launched from the MIPS site at 0053 UTC, 7 min after the GF passage (Fig. 21), sampled the trailing part of this wave region at 0100 UTC (assuming a sounding ascent rate of 5 m s−1). This sounding reveals the shallow depth of the cold outflow trailing the gust front (consistent with the SNR patterns in Fig. 19a), a residual layer between 0.4 and 1.6 km, a stable layer centered near 1.8 km, and a second mixed layer between 2 and 3 km. Unfortunately, this sounding terminated prematurely near 670 hPa. It does suggest stable layers near 800 and 680 hPa. A stable layer is also implied by the layer of enhanced 915-MHz SNR centered near 3.0 km between 0035 and 0055 UTC, and then extending up to 3.5 km by 0100 UTC (Fig. 19a). The main perturbation of ACB1 is located within this upper mixed layer. Small-scale variability in the sounding profiles of θ and rυ corroborate the large turbulent eddies inferred within this layer from the 915 MHz SNR and w, and ceilometer cloud-base measurements. Additional validation of relatively intense turbulence within ACB1 is indicated by high values of spectrum width (2.5–5.0 m s−1, not shown) measured by the 915-MHz profiler, and by clear air return (between −10 and 0 dBZ) measured near this level (3.5° elevation angle) by the KFSD radar (not shown).

These observations indicate a radical change in the structure of ACB1 from that produced by the collision of B1 and B2. Figure 14a indicates that ACB1 exhibited low-level convergence immediately after the collision at 2355 UTC, about 1 h earlier and 40 km north of the MIPS observations. It seems plausible that the vigorous gust front lifted ACB1 above the surface-based stable layer produced by cool air behind B2 and subsequent cloud shading. Based on the 915 measurements, the gust front head was about 1.3 km deep. If we assume that (i) the gust front head represents the maximum vertical displacement, (ii) the potential temperature of the mixed layer two hours earlier is 307 K (as indicated in the sounding shown in Fig. 2), and (iii) the surface temperature prior to the gust front was 28°C (as measured by the MIPS), then adiabatic lifting of the hypothetical sounding (shown in Fig. 21) by a vertical distance of 1.3 km would nearly duplicate the measured sounding. Since the gust front moved faster than ACB1, it would have intersected and then lifted the bore circulation about 1.3 km A continued upward propagation to the next stable layer (e.g., the top of the mixed layer near 700 hPa) would have occurred because of the near-neutral stability of the intervening layer. Although the MGLASS sounding in Fig. 21 represents the atmosphere modified by the gust front and passage of the elevated ACB1, the measured profiles of T and Td (θ and rυ) are consistent with the observed preexisting conditions and structure of the gust front.

This secondary collision event (gust front collision with a bore, which was previously produced by a collision between two gust fronts) represents the first detailed observation of a secondary collision. Although the detailed MIPS observations provide a comprehensive picture of the effects of this secondary collision, a numerical simulation would be required to clarify the physics of this interaction and the evolution ACB1. In this case, the active “weather” associated with ACB1 assumed the form of moderate to intense turbulence at scales ranging from about 103 m (as directly measured by the 915-MHz velocity) to about 10 m (as indicated by large values of spectrum width).

8. Summary and conclusions

This study has presented kinematic details of primary and secondary boundary collisions utilizing WSR-88D and MIPS data. The primary collision occurred when two mature gust fronts (B1 and B2) moving in opposite directions intersected very near the radar. The kinematic features of each gust front exhibited a difference in the updraft and body structure, apparently produced by their contrasting interaction with the ambient wind profile. Gust front B2, an older and less dense gravity current when compared to B1, moved at a slower ground-relative speed against the ambient flow, and thereby exhibited vertically erect updrafts at low levels. Two horizontal vortices that were aligned parallel to the gust front formed in the upper part of the B2 gravity current. For this reason, B2 was classified as a gravity current/wave hybrid. As B2 moved against the ambient vertical wind shear (more favorable conditions for CI), it initiated numerous convective cells.

In contrast, the eastward-moving boundary B1 (about 2 h old at the time of collision) was significantly faster because of the SW flow in the upper part of the boundary layer. Although maximum updrafts of 6 m s−1 within B1 were about 1 m s−1 greater than those in B2, B1 did not initiate deep convection prior to the collision, yet its systematic interaction with HCR updraft–downdraft couplets resulted in the formation of numerous nonprecipitating cumulus congestus clouds. At low levels B1 exhibited a rear-to-front feeder flow, while B2 did not.

At the time of the collision, a 4-km-deep updraft widened to 4 km and intensified to 12 m s−1. The collision did not produce CI at the initial collision point; rather intense deep convection form along the collision line 30–40 km south and north of the collision boundary axis. Both the magnitude and the depths of updrafts within B1 and B2 were enhanced as the two boundaries approached and collided. Kinematic analyses (WSR-88D) indicate that the after-collision boundaries exhibited bore and solitary wave characteristics, consistent with previous observations (Kingsmill and Crook 2003). The westward-moving boundary ACB2 developed a solitary wave structure with two well-defined radar finelines. The eastward-moving boundary also was surface rooted and exhibited borelike characteristics. Despite the minor difference in density of each gravity current, each after-collision boundary was surface rooted.

A secondary collision event involved a vigorous, new eastward-moving GF produced by deep convection initiated by the primary collision, and the eastward-moving bore (ACB1) produced 1 h earlier by the primary collision. This collision fortuitously occurred very close to the MIPS. The gust-front passage was accompanied by a 14 m s−1 wind gust. It propagated about 2 times faster than the boundary ACB1 and then intersected it several minutes before each boundary passed over the MIPS. The MIPS observations indicated that the gust front lifted the bore (ACB1) to the residual layer, well above the developing surface-based stable layer. We postulate that the bore continued its upward propagation within the neutral layer until it encountered a stable layer near the 2–3 km AGL level. The MIPS 915-MHz wind profiler sampled a well defined wave feature within 2–3 km layer, accompanied by a 3 m s−1 updraft, followed by a more intense 5 m s−1 downdraft, and relatively intense turbulence (where a subsequent sounding showed a neutral stratification). Several weaker w oscillations followed this initial wave.

Thus, the elevated form of ACB1 exhibited the structure of an internal gravity wave train that somewhat resembled a surface-based bore. This appears to be the first documented case of a sequence of boundary collisions that produced an elevated intense gravity wave.

Acknowledgments

This research was supported by the National Science Foundation (NSF) under Grants ATM-0239889 and ATM-0533596. The NSF provided support for the BAMEX field study. We acknowledge assistance from NCAR personnel who played an important role in the planning and field campaign phases of BAMEX.

REFERENCES

  • Atkins, N. T., R. M. Wakimoto, and C. L. Ziegler, 1998: Observations of the finescale structure of a dryline during VORTEX 95. Mon. Wea. Rev., 126 , 525550.

    • Search Google Scholar
    • Export Citation
  • Benjamin, J. B., 1968: Gravity current and related phenomena. J. Fluid Mech., 31 , 209248.

  • Charba, J., 1974: Application of gravity current model to analysis of squall-line gust front. Mon. Wea. Rev., 102 , 140156.

  • Cheung, T. K., and C. G. Little, 1990: Meteorological tower, microbarograph array, and sodar observations of solitary-like waves in the nocturnal boundary layer. J. Atmos. Sci., 47 , 25162536.

    • Search Google Scholar
    • Export Citation
  • Clarke, R. H., 1984: Colliding sea breezes and the creation of internal atmospheric bore waves: Two-dimensional numerical studies. Aust. Meteor. Mag., 32 , 207226.

    • Search Google Scholar
    • Export Citation
  • Clarke, R. H., R. K. Smith, and D. G. Reid, 1981: The morning glory of the Gulf of Carpentaria: An atmospheric undular bore. Mon. Wea. Rev., 109 , 17261750.

    • Search Google Scholar
    • Export Citation
  • Coleman, T. A., K. R. Knupp, and D. Herzmann, 2009: The spectacular undular bore in Iowa on 2 October 2007. Mon. Wea. Rev., 137 , 495503.

    • Search Google Scholar
    • Export Citation
  • Crook, N. A., 1986: The effect of ambient stratification and moisture on the motion of atmospheric undular bores. J. Atmos. Sci., 43 , 171181.

    • Search Google Scholar
    • Export Citation
  • Crook, N. A., 1988: Trapping of low-level internal gravity waves. J. Atmos. Sci., 45 , 15331541.

  • Crook, N. A., and M. J. Miller, 1985: A numerical and analytical study of atmospheric undular bores. Quart. J. Roy. Meteor. Soc., 111 , 225242.

    • Search Google Scholar
    • Export Citation
  • Davis, C., and Coauthors, 2004: The bow echo and MCV experiment: Observations and opportunities. Bull. Amer. Meteor. Soc., 85 , 10751093.

    • Search Google Scholar
    • Export Citation
  • Doviak, R. J., and R. Ge, 1984: An atmospheric solitary gust observed with a Doppler radar, a tall tower and a surface network. J. Atmos. Sci., 41 , 25592573.

    • Search Google Scholar
    • Export Citation
  • Doviak, R. J., S. Chen, and D. R. Christie, 1991: A thunderstorm-generated solitary wave observation compared with theory for nonlinear waves in a sheared atmosphere. J. Atmos. Sci., 48 , 87111.

    • Search Google Scholar
    • Export Citation
  • Droegemeier, K. K., and R. B. Wilhelmson, 1987: Numerical simulation of thunderstorm outflow dynamics. Part I: Outflow sensitivity experiments and turbulence dynamics. J. Atmos. Sci., 44 , 11801210.

    • Search Google Scholar
    • Export Citation
  • Fankhauser, J. C., N. A. Crook, J. Tuttle, L. J. Miller, and C. G. Wade, 1995: Initiation of deep convection along boundary layer convergence lines in a semitropical environment. Mon. Wea. Rev., 123 , 291313.

    • Search Google Scholar
    • Export Citation
  • Fovell, R. G., and P. S. Dailey, 2001: Numerical simulation of the interaction between the sea-breeze front and horizontal convective rolls. Part II: Alongshore ambient flow. Mon. Wea. Rev., 129 , 20572072.

    • Search Google Scholar
    • Export Citation
  • Frank, P. J., and P. A. Kucera, 2003: Radar characteristics of convection along colliding outflow boundaries observed during CRYSTAL-FACE. Preprints, 31st Int. Conf. on Radar Meteorology, Seattle, WA, Amer. Meteor. Soc., 12A.8.

    • Search Google Scholar
    • Export Citation
  • Fulton, R., D. S. Zrnic, and R. J. Doviak, 1990: Initiation of a solitary wave family in the demise of a nocturnal thunderstorm density current. J. Atmos. Sci., 47 , 319337.

    • Search Google Scholar
    • Export Citation
  • Intrieri, J. M., A. J. Bedard Jr., and R. M. Hardesty, 1990: Details of colliding thunderstorm outflows as observed by Doppler lidar. J. Atmos. Sci., 47 , 10811098.

    • Search Google Scholar
    • Export Citation
  • Karan, H., 2007: Thermodynamic and kinematic characteristics of low-level convergence zones observed by the mobile integrated profiling system. Ph.D. thesis. University of Alabama, 20 pp.

  • Karan, H., and K. R. Knupp, 2006: Mobile integrated profiler system (MIPS) observations of low-level convergent boundaries during IHOP. Mon. Wea. Rev., 134 , 92112.

    • Search Google Scholar
    • Export Citation
  • Kingsmill, D. E., 1995: Convection initiation associated with a sea-breeze front, a gust front, and their collision. Mon. Wea. Rev., 123 , 29132933.

    • Search Google Scholar
    • Export Citation
  • Kingsmill, D. E., and N. A. Crook, 2003: An observational study of atmospheric bore formation from colliding density currents. Mon. Wea. Rev., 131 , 29853002.

    • Search Google Scholar
    • Export Citation
  • Knupp, K. R., 2006: Observational analysis of a gust front to bore to solitary wave transition within an evolving nocturnal boundary layer. J. Atmos. Sci., 63 , 20162035.

    • Search Google Scholar
    • Export Citation
  • Koch, S. E., and W. L. Clark, 1999: A nonclassical cold front observed during COPS-91: Frontal structure and the process of severe storm initiation. J. Atmos. Sci., 56 , 28622890.

    • Search Google Scholar
    • Export Citation
  • Koch, S. E., P. B. Dorian, R. Ferrare, S. H. Melfi, W. C. Skillman, and D. Whiteman, 1991: Structure of an internal bore and dissipating gravity current as revealed by Raman lidar. Mon. Wea. Rev., 119 , 857887.

    • Search Google Scholar
    • Export Citation
  • Koch, S. E., C. Flamant, J. W. Wilson, B. M. Gentry, and B. D. Jamison, 2008: An atmospheric soliton observed with Doppler radar, differential absorption lidar, and a molecular Doppler lidar. J. Atmos. Oceanic Technol., 25 , 12671287.

    • Search Google Scholar
    • Export Citation
  • Mahapatra, P. R., R. J. Doviak, and D. S. Zrnić, 1991: Multisensor observation of an atmospheric undular bore. Bull. Amer. Meteor. Soc., 72 , 14681480.

    • Search Google Scholar
    • Export Citation
  • Mohr, C. G., L. J. Miller, R. L. Vaughn, and H. W. Frank, 1986: The merger of mesoscale data sets into a common Cartesian format for efficient and systemic analysis. J. Atmos. Oceanic Technol., 3 , 143161.

    • Search Google Scholar
    • Export Citation
  • Mueller, C. K., and R. E. Carbone, 1987: Dynamics of a thunderstorm outflow. J. Atmos. Sci., 44 , 18791898.

  • Oye, R., C. Mueller, and S. Smith, 1995: Software for radar translation, visualization, editing, and interpolation. Preprints, 27th Conf. on Radar Meteorology, Vail, CO, Amer. Meteor. Soc., 359–361.

    • Search Google Scholar
    • Export Citation
  • Ralph, F. M., C. Mazaudier, M. Crochet, and S. V. Venkateswaran, 1993: Doppler sodar and radar wind profiler observations of gravity-wave activity associated with a gravity current. Mon. Wea. Rev., 121 , 444463.

    • Search Google Scholar
    • Export Citation
  • Rottman, J. W., and J. E. Simpson, 1989: The formation of internal bores in the atmosphere: A laboratory model. Quart. J. Roy. Meteor. Soc., 115 , 941963.

    • Search Google Scholar
    • Export Citation
  • Seitter, K. L., 1986: A numerical study of atmospheric density current motion including the effects of condensation. J. Atmos. Sci., 43 , 30683076.

    • Search Google Scholar
    • Export Citation
  • Simpson, J. E., 1997: Gravity Currents in the Environment and the Laboratory. 2nd ed. Cambridge University Press, 244 pp.

  • Simpson, J. E., and R. E. Britter, 1980: A laboratory model of an atmospheric mesofront. Quart. J. Roy. Meteor. Soc., 106 , 485500.

  • Wakimoto, R. M., 1982: The life cycle of thunderstorm gust fronts as viewed with Doppler radar and rawinsonde data. Mon. Wea. Rev., 110 , 10601082.

    • Search Google Scholar
    • Export Citation
  • Wilson, J. W., and W. E. Schreiber, 1986: Initiation of convective storms at radar-observed boundary-layer convergence lines. Mon. Wea. Rev., 114 , 25162536.

    • Search Google Scholar
    • Export Citation

Fig. 1.
Fig. 1.

Upper air and surface observations on 16 Jun 2003. Solid and dashed contours depict isoheights/isobars and isobars/isotherms, respectively. The box shows the study area containing GBOS and KFSD WSR-88D radar. The line beginning at KFSD (K) and extending to the east represents 95 km distance where at the end convection initiation associated with synoptic forcing and BL circulations occurred. Soundings shown in Figs. 2, 7, and 21 were released from the MGLASS locations (G1 and G2) and the MIPS location (M).

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 2.
Fig. 2.

Skew T–logp diagram acquired from MGLASS (G2) sounding at 2235 UTC. See Figs. 1 and 4b for location.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 3.
Fig. 3.

Time series of KFSD VAD wind profile on 16 Jun 2003. Full and half barbs represent 10 and 5 kt, respectively. The LCL height from MGLASS sounding (Fig. 2) is annotated.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 4.
Fig. 4.

(a) Sioux Falls, SD, WSR-88D radar reflectivity factor at 1 km AGL at 2235 UTC. The figure also depicts the eastward-moving boundary (B1), the slow westward-moving boundary (B2), and radar finelines indicative of boundary layer roll circulations, varying from 0 to 10 dBZ, (b) Geostationary Operational Environmental Satellite-12 (GOES-12) visible image taken at 2201 UTC. Locations of MGLASS sounding units (G1 and G2), KFSD radar site, and MIPS are annotated.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 5.
Fig. 5.

Radar finelines of boundary B1 in dashed (boundary 2 in solid) moving eastward (westward). Distance between each ring is 15 km. Boundary collision time is about 2345 UTC (1845 CST). The location of MGLASS-1 shown with a dashed line is about 110 km from KFSD.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 6.
Fig. 6.

Radar reflectivity factor (increasing light blue to red color with maximum values of 55–60 dBZe) overlaid on surface observations. The oval and parallelogram depict regions having different CI modes. The dashed line (also drawn in Fig. 1b) represents 95 km distance. The dark arrow indicates the first cell formation associated with a HCR, and the white arrow shows the first cell initiated by synoptic forcing. Instrument and boundary locations (B1, B2) are also annotated.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 7.
Fig. 7.

MGLASS 1 sounding at 2201 UTC on 16 Jun 2003 (see Fig. 4b for location G1). The sounding was released behind B1. The letter h indicates the height of the density current, and d marks the height of the coldest temperature; θυ−cold and θυ−env represents average virtual potential temperatures of cold outflow and environmental air over the depth of h = 740 m.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 8.
Fig. 8.

Surface observations with 1-min time resolution acquired at Mitchell, SD, ASOS (see Fig. 5 for the location); (upper) virtual temperature (Tυ, solid), mixing ratio (rυ, dotted), and pressure (p, dashed) variations. (lower) Wind speed (spd, dotted), and wind direction (dir, solid). The passage of B1 occurred during the 2137–2144 UTC time period.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 9.
Fig. 9.

Sioux Falls, SD, (KFSD) WSR-88D reflectivity factor on an xy plane at 0.5 km AGL at (a) 2330, (b) 2345 UTC 16 Jun, and (c) 0030 and (d) 0055 UTC 17 Jun 2003. Boundaries defined in the text are labeled. Here, S refers to an intense storm, and M represents the MIPS location.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 10.
Fig. 10.

Vertical structures of B1 and B2. The xz cross sections pass through the radar location (y = 0) as shown in Fig. 9; CB1 and CB2 are propagation speeds of B1 and B2, respectively. (a) Eastward-moving boundary B1. Arrows depict the u–w ground relative flow. Shaded contours represent B1-relative airflow normal to B1. (b) Westward-moving boundary, B2, and u–w ground relative flow (arrows). Negative shaded values (light gray) indicate an easterly boundary-relative flow. The solid contours in (a) and (b) are reflectivity factor in dBZ.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 11.
Fig. 11.

Ground-relative flow within the xz plane passing through y = 0 at (a) 2320, (b) 2330, and (c) 2335 UTC. Reflectivity factor contours are solid lines. Thicker reflectivity contours depict values greater than 12 dBZ. The KFSD radar is located at the origin. The vertical axis is stretched by a factor of two. Refer to Fig. 9 for relative location. The wind profiles on the right were obtained from G2 sounding (Fig. 2) and the KFSD VAD analysis at 2330 UTC between 0.6 and 3 km MSL.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 12.
Fig. 12.

Surface observations (School Net) with 1-min time resolution from Sioux Falls—Pavilion, SD, between 2200 UTC 16 Jun and 0200 UTC 17 Jun 2003. Vertical dashed lines indicate passage of boundaries B2, ACB1, and GF.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 13.
Fig. 13.

Same as Fig. 11, but during and shortly after the collision on 16 Jun 2003.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 14.
Fig. 14.

Same as Fig. 11, but at (a) 2355 16 Jun 2003, and (b) 0000 and (c) 0015 UTC 17 Jun 2003.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 15.
Fig. 15.

(a) Radial velocity at 0.5 km AGL. Dashed line depicts the cross section taken in Fig. 16 along the A–B (Fig. 9d), which has 20° angle from west. Negative values are toward the radar, outbound winds are positive. Arrows depict updraft–downdraft couplets associated with ACB2 and its secondary circulation; (b) Z reflectivity factor at 0.5 km AGL.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 16.
Fig. 16.

Same as Fig. 11, but for boundary ACB2 along the dashed line drawn in Figs. 9d and 15 at 0055 UTC 17 Jun 2003. Thick line depicts the kinematic boundary ACB2 associated with a solitary wave with wavelength λ of 7.2 km and height h1 of 1.4 km.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 17.
Fig. 17.

Radar reflectivity factor (Z) at 0.5 km AGL at (a) 0035 and (b) 0045 UTC 17 Jun 2003. The KFSD radar is located at x = 0, y = 0. The circle depicts the MIPS location.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 18.
Fig. 18.

MIPS surface observations between 0037 and 0105 UTC 17 Jun 2003. (a) Time series of virtual potential temperature, mixing ratio, and pressure (gray), (b) wind speed and wind direction (gray). The vertical dashed lines depict the timing of the updrafts measured by the 915-MHz wind profiler ahead of the GF. The arrow in (a) indicates in the relative dynamic pressure maxima, and in (b) the sounding launch time.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 19.
Fig. 19.

Time–height section of 915-MHz profiler vertical beam measurements (a) SNR and (b) w acquired at 1-min intervals. (b) Values in black and white colors represent downdrafts exceeding 4 m s−1. (a),(b) Ceilometer-derived cloud-base height (black) is shown as solid dots. (c) Surface pressure time series. (d) Time–height section of ceilometer two-way attenuated backscatter profile. Values exceeding the color bar are in black. The inset in upper-right corner displays backscatter values between 0043 and 0053 UTC from surface to 0.7 km. (e) Time–height observations of water vapor mixing ratio acquired from the Microwave Profiling Radar (MPR). Measurements prior to 0051 UTC are not accurate because of warm up time. Mixing ratio contours are drawn for every 1 g kg−1 starting at 6 g kg−1. (f) Time series of mean vertical velocity acquired from 915-MHz profiler at 1 (solid) and 2.6 km (dotted) AGL. The CBZ passage occurred between 0043 and 0049 UTC. The solid vertical line represents the time of the sounding (0053 UTC) shown in Fig. 21.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 20.
Fig. 20.

(a) Time–height section of horizontal winds acquired by the 915-MHz profiler between 0030 and 0130 UTC. The solid gray line outlines the approximate gust frontal head region, and the dashed line defines the approximate top of the trailing density current. (b) The 2-min average winds derived from the Doppler sodar. The solid vertical line represents the time of the sounding (0053 UTC) shown in Fig. 21. Half and full barbs represent 2.5 and 5 m s−1 winds, respectively.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

Fig. 21.
Fig. 21.

Skew T–logp diagram acquired from the sounding released by the MIPS crew at 0053 UTC after the gust frontal passage. Mixing ratio lines are depicted in dotted lines slanted from left to right; θυ–cold is the average virtual potential temperature over the depth of 400 m; Tυ = 29°C is the virtual temperature prior to the GF passage.

Citation: Monthly Weather Review 137, 7; 10.1175/2008MWR2763.1

1

The average direction of the ambient winds within the 3 km depth was approximately from the southwest. Therefore, B2 was moving against the ambient flow.

2

The north–south component of the flow (υ) does not vary (i.e., the horizontal convergence is due to ∂u/∂x). See section 2 for details and justification of the 2D assumption.

3

Perhaps B2 could be interpreted as an incipient undular bore in view of the permanent upward displacement of air parcels that enter from the west, rise up to 1–1.5 km within the main updraft, and then continue moving eastward at this level. Note that such westerly flow does not exist west of B1 (Fig. 13a).

4

For incompressible flow the dynamic pressure is given by pdyn = 1/2ρ(υ22υ12), where, υ2 and υ1 are speeds of gust frontal flow and flow within the stagnation region perpendicular to the gust front. Inserting values of υ2 (9.8 m s−1) and υ1 (1.6 m s−1) yields a dynamic pressure of 0.38 mb, in close agreement with observed nonhydrostatic pressure increase of 0.36 hPa.

Save