+ All Categories
Home > Documents > UNIVERSITY OF OKLAHOMA GRADUATE COLLEGE …broberts/thesis/thesis_brett-roberts.pdf · 2012. 8....

UNIVERSITY OF OKLAHOMA GRADUATE COLLEGE …broberts/thesis/thesis_brett-roberts.pdf · 2012. 8....

Date post: 26-Jan-2021
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
104
UNIVERSITY OF OKLAHOMA GRADUATE COLLEGE TORNADOGENESIS IN HIGH-RESOLUTION IDEALIZED NUMERICAL SIMULATIONS A THESIS SUBMITTED TO THE GRADUATE FACULTY in partial fulfillment of the requirements for the Degree of MASTER OF SCIENCE IN METEOROLOGY By BRETT JULIAN ROBERTS Norman, Oklahoma 2012
Transcript
  • UNIVERSITY OF OKLAHOMA

    GRADUATE COLLEGE

    TORNADOGENESIS IN HIGH-RESOLUTION IDEALIZED NUMERICAL

    SIMULATIONS

    A THESIS

    SUBMITTED TO THE GRADUATE FACULTY

    in partial fulfillment of the requirements for the

    Degree of

    MASTER OF SCIENCE IN METEOROLOGY

    By

    BRETT JULIAN ROBERTS

    Norman, Oklahoma

    2012

  • TORNADOGENESIS IN HIGH-RESOLUTION IDEALIZED NUMERICAL

    SIMULATIONS

    A THESIS APPROVED FOR THE

    SCHOOL OF METEOROLOGY

    BY

    ______________________________

    Dr. Ming Xue, Chair

    ______________________________

    Dr. Howard Bluestein

    ______________________________

    Dr. Alan Shapiro

  • © Copyright by BRETT ROBERTS 2012

    All Rights Reserved.

  • iv

    Acknowledgements

    Financial support for this research came from NSF grant AGS-0802888 and an

    AMS Graduate Fellowship. The Extreme Science and Engineering Discovery

    Environment (XSEDE) provided supercomputing resources for this project, including

    systems administered by the National Institute for Computational Sciences (NICS) and

    Pittsburg Supercomputing Center (PSC).

    I am grateful to everyone who made this research and degree possible. Thanks

    are in order for Dr. Ming Xue, Dr. Howard Bluestein, and Dr. Alan Shapiro, who served

    on my committee. Dr. Xue, my primary advisor, obtained the funding for this project

    and provided much-needed perspective and experience through all its trials and

    tribulations. Drs. Bluestein and Shapiro reviewed this thesis and helped immensely in

    improving its quality through their insightful questions and comments. In addition,

    discussions with several of my fellow students were of great help as I attempted to

    analyze and make sense of my results. Alex Schenkman, in particular, spared significant

    time in assisting my progress and collaborating on areas of overlap between our

    respective projects. Daniel Betten also provided invaluable insight into my results, as

    did Dan Dawson, who in addition allowed me to use his code as the basis for many of

    the plots included in this thesis.

    I appreciate Scott Hill and Yunheng Wang of CAPS for helping to ensure the

    many computing resources I employed for this work ran as smoothly as is possible in

    the world of supercomputer-based modeling. Eileen Hasselwander and Debra Farmer in

    the CAPS office, along with Celia Jones and Marcia Pallutto in the School of

    Meteorology office, are thanked for their assistance in matters of scheduling and

  • v

    making sure the considerable paperwork entailed in the degree-seeking process reached

    its proper destination! Finally, I would be remiss not to thank all of my family and

    friends for their companionship and moral support as I worked toward this degree. My

    parents, in particular, have provided much encouragement and assistance (financial and

    otherwise!) over the years, and I would not be where I am today without it.

  • vi

    Table of Contents

    Acknowledgements ........................................................................................................ iv

    List of Tables ................................................................................................................ viii

    List of Figures ................................................................................................................ ix

    Abstract ......................................................................................................................... xii

    Chapter 1: Introduction and Organization .................................................................. 1

    Chapter 2: Background ................................................................................................. 3

    1. Initial development of a mid-level mesocyclone ............................................ 3

    2. Subsequent development of a near-ground mesocyclone .............................. 4

    a. Low-level vertical shear of the horizontal wind ....................................... 5

    b. Baroclinic generation of horizontal vorticity ........................................... 6

    c. Role of mid-tropospheric flow ................................................................. 7

    d. Dynamic pipe effect ................................................................................. 8

    e. Frictional effects ....................................................................................... 9

    3. Tornadogenesis ............................................................................................. 10

    Chapter 3: Methodology and Motivation ................................................................... 13

    1. Model configuration ..................................................................................... 13

    2. Experiments .................................................................................................. 15

    a. Control experiment ................................................................................. 15

    b. Microphysics experiments ...................................................................... 17

    c. Low-level shear experiments .................................................................. 20

    3. Definitions .................................................................................................... 26

    a. Mesocyclone ........................................................................................... 26

  • vii

    b. Tornado ................................................................................................... 26

    Chapter 4: Analysis of Control Simulation ................................................................ 28

    1. Storm morphology ........................................................................................ 28

    2. Low-level mesocyclone intensification and tornadogenesis ........................ 33

    a. Evolution of model fields ....................................................................... 33

    b. Trajectory analysis .................................................................................. 42

    3. Summary and Conclusions ........................................................................... 54

    Chapter 5: Sensitivity to Low-Level Shear ................................................................ 56

    1. Comparison of mesocyclone intensity and tornadogenesis timing .............. 56

    2. Qualitative comparison and storm morphology ........................................... 60

    3. Summary and Conclusions ........................................................................... 68

    Chapter 6: Sensitivity to Microphysics Parameterization ........................................ 70

    1. Comparison of low-level mesocyclone intensity and tornadogenesis timing

    ...................................................................................................................... 70

    2. Qualitative comparison and storm morphology ........................................... 74

    3. Summary and Conclusions ........................................................................... 78

    Chapter 7: Comments on High-Resolution Model Trajectory Calculations .......... 80

    1. Accuracy in the presence of noise ................................................................ 80

    2. The use of Eularian temporal interpolation for trajectory calculations ........ 82

    References ..................................................................................................................... 86

    Appendix A: Sounding Wind Data for Shear Experiments ..................................... 90

    Appendix B: Numerical Methods for Trajectory Calculations ................................ 91

  • viii

    List of Tables

    Table 1. ARPS model configuration (universal to all experiments). ............................. 14

    Table 2. Summary of soundings and intercept values for all experiments. .................... 23

    Table 3. Low-level vertical wind shear magnitudes for all shear experiments. ............. 24

    Table 4. Trajectory experiments for temporal interpolation comparison. ...................... 83

    Table 5. Median initial position error for 300 s backward trajectories. ......................... 84

    Table 6. Mean initial position error for 300 s backward trajectories. ............................ 84

    Table 7. Zonal (u) wind for sounding points below 2 km AGL, in m s-1

    . ...................... 90

    Table 8. Meridional (v) wind for sounding points below 2 km AGL, in m s-1

    . ............. 90

  • ix

    List of Figures

    Figure 1. Sounding “may20” utilized for the control experiment (CTRL). ................... 17

    Figure 2. Graphical depiction of parameter space covered by all experiments. ............. 23

    Figure 3. Zonal (u) wind component for the six initial soundings. ................................ 24

    Figure 4. Meridional (v) wind component for the six initial soundings. ........................ 25

    Figure 5. Average zonal (u) and meridional (v) wind components for strongly-tornadic

    (solid), weakly-tornadic (dashed), and nontornadic (dotted) supercell environments as

    derived from 413 RUC proximity soundings. Adapted from Markowski et al. (2003). 25

    Figure 6. Perturbation potential temperature (shading; K) and 1 g kg-1

    rainwater mixing

    ratio contour at 10 m AGL for (a) 3000 s, (b) 4800 s, (c) 7200 s, and (d) 10200 s. ...... 30

    Figure 7. As in Fig. 5, but for vertical velocity (shading; m s-1

    ) at 4265 m AGL. ......... 31

    Figure 8. Close-up (moving) view of low-level mesocyclone at 10 m AGL. Perturbation

    potential temperature (shading; K) and 0.1 s-1

    vertical vorticity contour (orange solid)

    are plotted at (a) 9000 s, (b) 9420 s, (c) 9600 s, and (d) 9900 s. Gust fronts are indicated

    by solid red (FFGF), dark blue (RFGF), and purple (occluded) curves. Black “T”

    indicates tornado position. .............................................................................................. 34

    Figure 9. Zoomed (moving) view of the tornado in CTRL. (a-d) Horizontal wind speed

    (shading; m s-1

    ) and perturbation pressure at 3 hPa increments below -3 hPa (contour),

    at 10 m AGL. (e-h) Vertical velocity (shading; m s-1

    ) and p’ (contour; -5 hPa) at 1039 m

    AGL. Pretornadic vortices are denoted by “PV1” and “PV2” in (a). ............................. 36

    Figure 10. Vertical cross-section through (a) pretornadic vortex PV1 at 9180 s, and (b-

    d) the tornado at 9480, 9600, and 9720 s. Vertical velocity (shading; m s-1

    ) is shown,

    along with positive vertical vorticity contours (s-1

    ) at 0.1 s-1

    intervals starting at 0.1 s-1

    .

    The horizontal endpoints of each slice are indicated as (xs, ys)...................................... 39

    Figure 11. Conceptual model of a two-cell vortex associated with subcritical flow.

    Adapted from Trapp (2000). ........................................................................................... 39

    Figure 12. Time-height cross-section of (a) maximum vertical velocity, (b) maximum

    vertical vorticity, and (c) minimum perturbation pressure over the model domain from

    7200-12600 s. ................................................................................................................. 41

    Figure 13. Parcel trajectories initialized in the vortex at (a) 9480 s, (b) 9600 s, and (c)

    9780 s; and in the low-level mesocyclone center at (d) 10080 s. Parcels were initialized

    at 50 m AGL and integrated backward in time for 480 s. Parcel height is indicated by

    color, providing a pseudo-3D visualization. ................................................................... 44

  • x

    Figure 14. Simulated reflectivity at 1 km AGL at (a) 9240 s, (b) 9360 s, (c) 9480 s, and

    (d) 9600 s. Black circles denote the leading edge of the RFD surge at this height. ....... 46

    Figure 15. Horizontal cross-sections of simulated reflectivity following a representative

    tornado-entering parcel. The height of each slice is the parcel height at the indicated

    time. The black curve is the projection of the parcel path into the xy-plane, while the

    large dot is its current location. ...................................................................................... 47

    Figure 16. Horizontal projection of the trajectory path for the representative parcel.

    Color indicates parcel height, allowing for a pseudo-3D visualization. The remaining 20

    parcels which were also initialized at 9480 s are shown in gray. Stars denote original

    parcel locations (at 9000 s). ............................................................................................ 49

    Figure 17. Diagnostic quantities calculated along the representative trajectory from

    9000-9480 s. (a) Vertical and horizontal-streamwise vorticity components, (b) source

    terms for horizontal-streamwise vorticity, and (c) parcel height.................................... 50

    Figure 18. Horizontal cross-sections of equivalent potential temperature (shading; K)

    for 9240 s and 9300 s. The projection of the parcel path into the xy-plane is indicated by

    the black curve, while the large black dot is its current location. The height plotted at

    each time is the parcel height at that time. ..................................................................... 53

    Figure 19. Timeseries of (a) minimum perturbation pressure, and (b) maximum surface

    wind speed, for all low-level shear experiments over the period 7200-12600 s. ........... 58

    Figure 20. Time-height cross sections of minimum perturbation pressure for all low-

    level shear experiments over the period 7200-12600 s. ................................................. 59

    Figure 21. Timeline of tornado occurrence for the low-level shear experiments. ......... 60

    Figure 22. Perturbation potential temperature (shading; K) and 1 g kg-1

    rainwater

    mixing ratio contour at 10 m AGL and 8400 s for all low-level shear experiments. ..... 61

    Figure 23. Perturbation potential temperature (shading; K) at the time of peak tornado

    intensity for all low-level shear experiments containing a tornado. The RFGF is denoted

    by a dark blue curve, while the occluded portion of the gust front is purple. Red “T”

    indicates tornado position. Note that for IS2, the first of two tornadoes is shown. The

    plotted time varies by experiment and is indicated below the experiment name. .......... 62

    Figure 24. Backward parcel trajectories initialized in tornadoes from four low-level

    shear experiments; see Section 4.2b for complete methodology description. Trajectories

    were integrated backward in time for 480 s. For each experiment, the initialization time

    (from which backward integration was performed) is listed below the experiment name.

    ........................................................................................................................................ 64

    Figure 25. Vertical cross-section through the strong tornado in IS2 at (a) 10740 s and

    (b) 10920 s. Vertical velocity (shading; m s-1

    ) is plotted, along with contours for

    positive vertical vorticity at 0.1 s-1

    increments starting at 0.1 s-1

    . .................................. 66

  • xi

    Figure 26. Horizontal path for the representative parcel entering the strong tornado in

    IS2 at 10740 s. Parcel height is indicated by color. The remaining 20 parcels calculated

    for this initial time are plotted in gray. Stars indicate original positions after backward

    integration. ...................................................................................................................... 66

    Figure 27. Diagnostic calculations of (a) vertical and horizontal streamwise vorticity,

    (b) horizontal streamwise vorticity source terms, and (c) parcel height, for the

    representative parcel entering the strong tornado in IS2. ............................................... 67

    Figure 28. Timeseries of (a) minimum perturbation pressure, and (b) maximum surface

    wind speed, for all microphysics experiments over the period 7200-12600 s. .............. 72

    Figure 29. Time-height cross sections of minimum perturbation pressure for all

    microphysics experiments over the period 7200-12600 s. ............................................. 73

    Figure 30. Timeline of tornado occurrence for the microphysics experiments. ............. 74

    Figure 31. Perturbation potential temperature (shading; K) and 1 g kg-1

    rainwater

    mixing ratio (solid contour) at 8400 s for all microphysics experiments. ...................... 77

    Figure 32. Perturbation potential temperature (shading; K) and wind vectors at 10 m

    AGL for (a) R86, (b) R45, (c) CTRL, and (d) R45H43. Tornado centers at 10 m AGL

    are denoted by “T,” mid-level mesocyclone centers at 2925 m AGL are denoted by

    “M,” and RFGF positions are highlighted by dark blue curves. Note that the position of

    the plotted subdomain within the model domain varies for each experiment. ............... 78

    Figure 33. Timeseries of Eulerian and Lagrangian horizontal streamwise vorticity for

    the representative parcel in Section 4.2b. ....................................................................... 81

    Figure 34. Timeseries of (a) Eulerian and Lagrangian horizontal streamwise vorticity,

    and (b) vertical velocity, for Parcel B. Red and orange ellipses denote two periods in

    which disagreement between Eularian and Lagrangian time tendency for streamwise

    vorticity is very large. ..................................................................................................... 81

  • xii

    Abstract

    For several decades, idealized numerical simulations wherein an initial thermal

    bubble induces deep moist convection have been an important tool for understanding

    supercells. Several such studies in the literature have utilized the 20 May 1977 “Del

    City storm” sounding for their initial environment, often employing nested grids to

    attain higher resolution over some spatiotemporal subdomain of interest. In this study,

    simulations are conducted at uniform 50 m horizontal resolution using this sounding in

    the Advanced Regional Prediction System (ARPS) with Lin ice microphysics. Under

    this configuration, the initial right-moving supercell fails to produce a tornado within

    the first 2 h of model integration. However, a second convective cell develops to its

    southwest, and an eventual merger of the two storms leads to tornadogenesis later in the

    simulation. The resulting tornado is a two-cell vortex with central downdraft throughout

    its life cycle, likely owing to the lack of surface friction in the model. Diagnostic

    calculations along a representative tornado-entering trajectory reveal that both

    baroclinic generation and tilting of background vorticity contribute in comparable

    measures to positive streamwise vorticity before the parcel turns upward near the

    vortex.

    In addition to the control simulation, two sets of sensitivity experiments are

    performed. In the first set, the rain and hail intercept parameter values are varied to test

    sensitivity to precipitation microphysics. In the second set, the distribution of vertical

    shear within the lowest 2 km AGL is varied around the original 20 May 1977 wind

    profile. In both cases, a clear temporal trend in overall storm evolution is identified:

    faster evolution for weaker 0-1 km AGL shear and for smaller intercept parameter

  • xiii

    values. However, the effect upon tornado occurrence, intensity and duration is nonlinear

    in both cases. Finally, some issues relating to trajectory accuracy in high-resolution

    supercell simulations are briefly explored.

  • 1

    Chapter 1: Introduction and Organization

    This study presents a suite of high-resolution idealized numerical simulations of

    supercell thunderstorms. There are two primary goals. The first is to understand in a

    general sense the storm evolution and how it leads to low-level mesocyclogenesis and

    tornadogenesis. The second is to explore the sensitivity of this evolution to a parameter

    space in both low-level shear distribution and microphysics parameterization. As

    detailed in Chapter 3, a key strength of this study is the use of a very high (50 m)

    uniform horizontal grid spacing. This allows the identification of tornadoes within the

    simulations, where earlier studies either could only identify tornado cyclones (owing to

    lower resolution) or relied on nested subdomains to resolve tornadoes themselves.

    This thesis is organized as follows. Chapter 2 presents a brief literature review

    of supercells, mesocyclogenesis and tornadogenesis. Chapter 3 introduces the

    experimental design and motivation for this work.

    Chapter 4 presents an analysis of the control simulation. First, the storm-scale

    evolution over the course of the simulation is described and illustrated. Then, the period

    in which the low-level mesocyclone intensifies and a tornado occurs is analyzed in

    more detail. Trajectories entering the low-level mesocyclone and tornado are examined,

    and a vorticity budget is calculated along one representative parcel.

    Chapter 5 compares the experiments in which low-level shear was varied from

    the control run. Similarly, Chapter 6 compares the microphysics-varying experiments.

    In both cases, temporal trends in storm evolution – particularly with respect to the

    development of a low-level mesocyclone – are identified within the parameter space.

    Furthermore, the relative tendency for storm features to become configured in a manner

  • 2

    favoring tornadogenesis over the parameter space is explored. In the case of the low-

    level shear experiments, one experiment produces a tornado that is explored in some

    detail due to important differences from the others.

    Chapter 7 adds a few comments on the use of trajectories as a diagnostic tool in

    model simulations, particularly at spatial resolutions considered high at the time of this

    study (Δx < 100 m). It is demonstrated that such trajectories may sometimes contain

    significant position errors when passing through fast, complex flow, as exemplified by

    the region near a tornado. Within the context of trajectory calculations, the value of

    employing temporal interpolation between model output times is also explored.

  • 3

    Chapter 2: Background

    Despite receiving considerable attention from the research community in recent

    decades, our understanding of the physical processes responsible for supercell

    tornadogenesis remains relatively poor. One need look no further for evidence of this

    inadequacy than in the operational setting, where forecasters continue to issue tornado

    warnings in a primarily reactive fashion upon detection by radar or spotters (Stensrud et

    al. 2009). Recent observational and numerical studies have made progress in supporting

    or refuting mechanisms hypothesized to be responsible for the generation of tornadic-

    strength vorticity within supercell thunderstorms, serving to guide current research

    towards the most likely candidates. Nevertheless, the roles and relative contributions of

    these candidates remain greatly in question.

    Numerical and observational work has identified three processes which typically

    occur sequentially leading up to supercell tornadoes: first, the development of a mid-

    level mesocyclone; second, the development of a low-level (or near-ground)

    mesocyclone; and finally, tornadogenesis itself. Each stage is reviewed below.

    1. Initial development of a mid-level mesocyclone

    Observations indicate that supercells first develop a mid-level mesocyclone

    several km above ground level (AGL), and only afterwards gain a near-ground

    mesocyclone at or below 1 km AGL (Davies-Jones et al. 2001). In his summary of the

    current state of tornadogenesis research, Davies-Jones (2006) suggested that the

    development of this near-ground mesocyclone is the least-understood process in the

    chain of events which leads to supercell tornadoes – as such, it will be the main focus of

    this review The preceding development of a mid-level mesocyclone is understood to

  • 4

    result from tilting of environmental horizontal streamwise vorticity into the vertical by a

    supercell’s updraft (Rotunno and Klemp 1985). Davies-Jones (1984) derived the

    theoretical correlation coefficient between perturbation vertical velocity (w’) and

    vertical vorticity (ζ’), showing that it varies directly with the proportion of

    environmental vorticity which is streamwise:

    where is the streamwise component of the environmental vorticity vector. The

    physical significance of this expression is that larger environmental streamwise vorticity

    implies a greater tendency for updrafts (positive w’) to rotate cyclonically (positive ζ’).

    This idea is supported by the finding by Droegemeier et al. (1993) that 0-3 km storm-

    relative environmental helicity (SREH), a vertical integral of environmental streamwise

    vorticity, is useful in discriminating between supercell and nonsupercell thunderstorms

    in a numerical model.

    2. Subsequent development of a near-ground mesocyclone

    Near-ground mesocyclones are unlikely to share the direct vorticity-tilting mode

    of formation, however. Such a mechanism would require near-ground parcels bearing

    horizontal streamwise vorticity to turn upward quite sharply in order to produce strong

    vertical vorticity close to the ground, and observations do not support the existence of

    such extreme horizontal gradients in near-ground vertical velocity prior to

    tornadogenesis (Bluestein 2007). Therefore, other sources for either creating or

    amplifying existing near-ground vertical vorticity must be identified.

  • 5

    a. Low-level vertical shear of the horizontal wind

    The robust correlation observed by forecasters and formally established by

    Rasmussen (2003) between high values of 0-1 km SREH and the likelihood that a

    supercell will produce a tornado seems curious, given that near-ground mesocyclones (a

    prerequisite for supercell tornadoes) probably do not form as a direct result of an

    updraft tilting horizontal streamwise vorticity into the vertical. Davies-Jones (2006)

    proposed that the base of a mid-level mesocyclone may be relatively low when 0-1 km

    SREH is high, perhaps encouraging or hastening the formation of a near-ground

    mesocyclone in some way. This idea is supported by Schumacher and Boustead (2011),

    who associated an increasing magnitude of 0-1 km shear during one outbreak with an

    abrupt increase in the number and strength of tornadoes. Another possibility is that

    horizontal streamwise vorticity associated with low-level environmental shear is

    amplified by stretching as parcels descend in the rear-flank downdraft (RFD), and upon

    reaching the surface these parcels exhibit high enough streamwise vorticity that only

    modest tilting and stretching near the ground by an updraft is necessary to instigate a

    near-ground mesocyclone.

    Mashiko et al. (2009), hereafter M09, numerically simulated a minisupercell

    associated with a typhoon’s outer rain band, and found through trajectory analysis that

    about half the parcels entering the storm’s intensifying low-level mesocyclone 18 s

    prior to tornadogenesis originated in the RFD. The vorticity budget for a representative

    parcel was analyzed, and stretching of already-substantial streamwise vorticity from the

    environmental shear was found to be the primary contributor to its large value after

    descending toward the near-ground mesocyclone center. While tropical minisupercells

  • 6

    typically form in environments of notably weaker buoyancy and stronger low-level

    shear than classic supercells, the findings of M09 lend credence to the more general

    idea that the RFD aids near-ground supercell mesocyclogenesis through barotropic

    downward transport of large environmental streamwise vorticity. Moreover, earlier

    numerical studies of classic supercells have demonstrated evidence that some parcels

    entering the low-level mesocyclone originate from the RFD and experience stretching

    of streamwise vorticity during descent (Adlerman et al. 1999).

    b. Baroclinic generation of horizontal vorticity

    Another hypothesized mechanism for near-ground mesocyclogenesis is the

    baroclinic generation of horizontal vorticity which is tilted into the vertical and

    stretched. Clearly, this explanation faces the same obstacle regarding the need for

    unrealistically sharp horizontal gradients in low-level vertical motion discussed in the

    preceding section. Proposed sources for baroclinic generation include the storm-scale

    forward-flank downdraft (FFD) gust front, as well as mesoscale and synoptic-scale

    outflow boundaries and fronts. The former was found to be of significant importance in

    the numerical simulations of Klemp and Rotunno (1983). However, observations during

    the VORTEX field experiment in 1994-95 revealed an unexpected lack of baroclinicity

    along the FFD gust fronts of many significantly tornadic supercells. In fact, Shabbott

    and Markowski (2006) analyzed mobile mesonet data from 12 supercells and found

    those whose FFD had large θ deficits at the surface (and therefore a relatively strong

    gradient of buoyancy along the gust front) were actually less likely to be tornadic than

    those with smaller θ deficits in the FFD. Moreover, Markowski et al. (2002) observed a

    similar pattern for the thermodynamics of the RFD. These studies and others have cast

  • 7

    significant doubt on the role of baroclinic streamwise vorticity generated at the storm

    scale in low-level rotation, prompting researchers to look elsewhere over the past

    decade.

    Markowski et al. (1998), hereafter M98, noted that a large proportion of the

    tornadoes observed during VORTEX occurred when supercells interacted with

    mesoscale boundaries, a phenomenon also frequently observed by operational

    meteorologists. As such, unlike the FFD buoyancy gradient, the causal role of this form

    of baroclinic vorticity generation in tornadogenesis cannot easily be dismissed.

    Reconciling this fact with the need for very abrupt upward turning of parcels traveling

    along these boundaries in order to produce a near-ground mesocyclone remains a

    challenge facing research meteorologists. It is possible that inflow parcels which travel

    along the baroclinic boundary and acquire enhanced streamwise vorticity enter the

    updraft, wrap around it cyclonically in the mid-level mesocyclone, and are later

    processed by the RFD, ultimately descending and converging into the low-level

    mesocyclone. Another potential factor noted by M98 is that updraft intensity may

    increase due to enhanced low-level convergence along mesoscale boundaries, which

    could at the very least increase vertical stretching of parcels in an incipient tornado

    given a pre-existing near-ground mesocyclone.

    c. Role of mid-tropospheric flow

    In addition to the generation of vorticity discussed in 3a, vertical shear of the

    horizontal wind likely influences the development of a low-level mesocyclone in other

    capacities. Brooks et al. (1994) found that the development and longevity of near-

    ground rotation in a supercell depends in part upon a complex balance between mid-

  • 8

    level storm-relative winds and low-level vertical shear (and resultant mid-level

    mesocyclone intensity). In cases with very strong mid-level winds, precipitation may be

    advected far downshear from the updraft, which is rich in streamwise vorticity. Hence,

    the precipitation-induced downdraft will form in a location well removed from a major

    vorticity source, unable to transport it downward efficiently for near-ground

    mesocyclogenesis. Conversely, weak mid-level winds can result in precipitation falling

    into and weakening the updraft, or at least falling close enough to the updraft that cold

    outflow undercuts it (and the near-ground mesocyclone) more quickly. An optimal

    balance between the two ensures that the RFD is close enough to the updraft to deliver

    vorticity-rich parcels, but not so close that updraft intensity is adversely affected.

    It is important to note that this finding implies the vertical wind profile

    throughout a large depth of the troposphere, not just the lowest few kilometers, is a

    control on the formation of low-level mesocyclones.

    d. Dynamic pipe effect

    A theoretical means by which an atmospheric vortex in cyclostrophic balance

    may build downward with time through the “suction” of vorticity-bearing air below was

    proposed by Leslie (1971), who termed it the dynamic pipe effect (DPE). In this

    situation, the pre-existing vortex acts much like a pipe, drawing up air from below

    through the vertical perturbation pressure gradient force (VPPGF) associated with

    rotationally-induced low pressure in the vortex. The horizontal convergence of inflow

    air at the bottom of the vortex would amplify any existing vertical vorticity through

    stretching, and if this effect were strong enough, downward propagation of the vortex

  • 9

    would occur. Once established, this process would continue at successively lower levels

    until the vortex reached the ground.

    While it might seem tempting to employ this explanation as the mechanism for

    near-ground mesocyclogenesis following the establishment of mid-level rotation,

    Davies-Jones et al. (2001) assert that mid-level mesocyclones are not in cyclostrophic

    balance, and are thus not resistant to parcel displacements in the radial direction. The

    consequence of this idea is that air may enter the mesocyclone through its sides,

    precluding the formation of a vacuum at its base (or top). Recent observations from

    mobile phased-array radars with high temporal data frequency have revealed further

    weaknesses in the DPE hypothesis. French (2012) found that despite the appearance of

    a descending TVS in five-minute WSR-88D radar data for the 5 June 2009 tornado in

    Goshen Co., WY, there actually was none in mobile phased array data. Instead, high-

    frequency velocity data revealed many short-lived mid-level TVS signatures which

    might easily be mistaken for one continuous TVS (and therefore a descending vortex

    with time) in the 88D data. As such, the DPE is unlikely to explain near-ground

    mesocyclogenesis, and other sources of enhanced low-level vorticity (such as those

    described in 3a-b) must be considered.

    e. Frictional effects

    Alongside baroclinic generation and tilting of environmental shear, another

    possible source of vorticity for low-level mesocyclones is that generated by surface

    friction. Although this factor has been remained somewhat unstudied relative to the

    former two, some recent work is beginning to shed light on it. In the observational

    realm, Markowski et al. (2012) examined vortex lines and material circuits derived from

  • 10

    dual-Doppler analyses of the 5 June 2009 tornadic supercell. When they were unable to

    reconcile a calculated increase in circulation about a tornado-enclosing circuit with the

    value expected from baroclinic generation, the possibility that surface friction

    accounted for the discrepancy was raised.

    Schenkman et al. (2012) demonstrated the critical importance of surface friction

    in the development of a tornado-like vortex within a mesoscale convective system

    (MCS). Specifically, the inclusion of friction in their simulations led to a horizontal

    rotor and associated strong low-level updraft which enhanced vortex stretching.

    Although the broader convective system which produced this vortex differs from a

    classic supercell in some important respects, the authors note the presence a low-level

    mesovortex which resembles a supercellular low-level mesocyclone in structure and

    spatiotemporal scale. Thus, their work shows convincingly that surface friction can

    strongly influence tornadogenesis in at least some circumstances. It must be

    emphasized, however, that its role as a vorticity source was found to be secondary to

    that of augmenting the low-level updraft and stretching existing vorticity.

    3. Tornadogenesis

    Following the establishment of a strong near-ground mesocyclone, locating the

    source of vorticity for tornadogenesis becomes much easier. Bluestein (2007)

    demonstrates through scale analysis that convergence of O(10-2

    s-1

    ) acting upon a low-

    level mesocyclone ~5 km in horizontal diameter could produce vorticity of tornado

    strength, O(1 s-1

    ), on a timescale of 1000 s. Because the specified magnitude of

    convergence is typical of values found beneath intense supercell updrafts, and the

    specified timescale is consistent with that of tornadoes, it is easily seen that

  • 11

    convergence/stretching of mesocyclonic vertical vorticity alone can account for

    tornadogenesis.

    The Tornado Vortex Signature (TVS), first described by Brown et al. (1978), is

    a radar signature indicative of strong azimuthal wind shear a few kilometers AGL and is

    associated with many tornadoes. Trapp and Davies-Jones (1997) distinguished between

    cases in which a TVS precedes tornadogenesis by at least 5 min. (mode I

    tornadogenesis) and cases in which a TVS appears nearly simultaneously with

    tornadogenesis (mode II). Mode I corresponds to a situation in which mid-level rotation

    is initially stronger than near-ground rotation, possibly necessitating the DPE in order to

    build the tornado vortex downwards to the ground. (Note that, unlike a mesocyclone,

    the tornado vortex is in cyclostrophic balance and resists parcel motion in the radial

    direction, so the DPE is plausible). Thus, in cases where low-level convergence beneath

    the updraft is initially insufficient to instigate a tornado, the DPE may provide the extra

    convergence needed to begin the process.

    Interestingly, Trapp et al. (1999) examined the TVS character associated with 52

    tornadoes and found roughly an equal number of descending and nondescending

    signatures. This conflicts with the assumption that supercell tornadoes are usually of the

    descending variety, since most tornadoes are supercellular; indeed, closer examination

    of the sample set by the authors revealed various supercell tornado cases with

    nondescending TVS behavior. This is a confirmation that the DPE is not always

    necessary for tornadogenesis, and more generally, that supercell tornadoes may not

    always build down from above. The idea that existing low-level convergence is

    sufficient in some instances also has implications for the role of mesoscale boundaries

  • 12

    in tornadogenesis described in 3b: the role of enhanced convergence along such

    boundaries must be carefully considered as an alternative or complement to their

    baroclinic vorticity generation in explaining why tornadoes are frequently associated

    with them.

  • 13

    Chapter 3: Methodology and Motivation

    1. Model configuration

    This study examines a set of numerical simulations conducted using the

    Advanced Regional Prediction System (ARPS), a three-dimensional non-hydrostatic

    model developed at the Center for Analysis and Prediction of Storms (Xue et al. 2000).

    Version 5.2.12 of the ARPS package was used for model integration, and the message-

    passing interface (MPI) system was employed by splitting the model domain into 1024

    equally-sized horizontal patches, with one CPU core assigned to each patch (Johnson et

    al. 1994).

    A summary of the model configuration for these simulations is found in Table 1.

    The horizontal resolution is uniform at 50 m; this is considered sufficient to resolve

    tornadoes (albeit not their detailed internal structure), which distinguishes this study

    from most work in the existing literature based upon the Del City sounding. The vertical

    resolution varies with height from 20 m near the ground to 400 m at the top of the

    domain (16 km AGL), and there are 83 vertical levels. Lateral boundary conditions are

    radiation, while top and bottom boundary conditions are free-slip rigid wall. A Rayleigh

    sponge layer is present from 12 km AGL upward in order to minimize the effects of

    waves reflecting off the top of the domain.

    All simulations herein are idealized, i.e., they are initialized in a horizontally-

    homogeneous manner from a single sounding and otherwise do not assimilate any real

    data, nor do they contain terrain at the model surface. The Coriolis force and surface

    physics (including friction) are neglected. Deep moist convection is induced by

    introducing a “bubble” with an initial positive perturbation in potential temperature of

  • 14

    magnitude 4 K. The bubble has a radius of 5x5x1.5 km, and within the 64x64x16 km

    model domain, is centered at (x, y, z) = (46, 28, 1.5) km. This position was determined

    through trial and error with the goal of keeping the entire storm of interest inside the

    domain after several hours of integration.

    Integration was allowed to proceed for 12600 s (3.5 h), with history files saved

    every 60 s. For some experiments, a period of interest was identified for which higher

    temporal resolution was desired, particularly for the purpose of trajectory analysis. In

    these cases, a restart file written during the initial run was used to re-initialize the model

    at or near the beginning of the period in question and write 3 s output. Due to the

    relatively small model time step and use of MPI, the model solution during this

    subsequent integration diverges from the original solution slightly with time, but the

    results were found to be qualitatively similar even at tornado scale.

    Parameter Value

    Domain size 64x64x16 km

    Grid size 1283x1283x83

    Horizontal resolution, Δx 50 m

    Vertical resolution, Δz 20 m ≤ Δz ≤ 400

    m Large time step, Δt 0.25 s

    Small time step, Δτ 0.08 s

    Microphysics parameterization Lin ice

    Turbulence parameterization 1.5-order TKE

    Mixing coefficient, horiz. (4th

    order) 20x10-4

    s-1

    Mixing coefficient, vert. (4th

    order) 10x10-4

    s-1

    Table 1. ARPS model configuration (universal to all experiments).

  • 15

    2. Experiments

    a. Control experiment

    A control run (CTRL) was initialized using the same 20 May 1977 sounding

    employed in numerous earlier supercell modeling studies, notably Klemp et al. (1981),

    Adlerman et al. (1999) (hereafter A99), and Noda and Niino (2010) (hereafter NN10).

    This sounding is a blend of observations from Ft. Sill and Elmore City, OK, during the

    evening of a tornado event in central Oklahoma, and features a classic Great Plains

    supercell environment with surface-based convective available potential energy

    (SBCAPE) of approximately 4000 J kg-1

    , effective storm-relative helicity (SRH) of 101

    m2 s

    -2, and a veering vertical wind profile which produces a hodograph shape

    resembling a semicircle. Figure 1 presents this sounding graphically. The horizontal

    resolution of 50 m is finer than any published study with this sounding to the author’s

    knowledge, with the closest (and most recent) contender being NN10, who used Δx =

    70 m. As with previous ARPS model studies, the sounding has been interpolated above

    the surface to 500 m vertical resolution, with state variables defined between 250 m and

    16250 m AGL. Note that upon model initialization the sounding data is then re-

    interpolated to the grid itself, which is stretched in the vertical. The 20 May 1977

    observed storm motion of (u,v) = (3,14) m s-1

    is subtracted from the wind data

    throughout the sounding. This results in a model domain that effectively follows the

    storm motion without the need to handle explicitly a moving domain. As such, any

    references to cardinal directions and speeds regarding motion hereafter must be

    interpreted in a quasi-storm-relative, rather than ground-relative, framework.

  • 16

    The second major difference between CTRL and earlier, comparable work is the

    microphysics treatment. Whereas A99 and NN10 both employed the Kessler warm rain

    parameterization, this simulation (and indeed all others herein) instead uses Lin ice,

    which is further detailed in the next subsection. Though this choice carries some

    computational expense, the inclusion of a hail species in particular should improve the

    accuracy of storm thermodynamics and cold pool evolution, an important factor for

    tornadogenesis. For CTRL, rain and hail intercept parameter values are 8x10-4

    m-4

    and

    4x10-4

    m-4

    , respectively. See Table 2 for a summary of the differences between all

    experiments. Figure 2 provides a graphical description of the parameter space covered

    by the experiments herein.

  • 17

    Figure 1. Sounding “may20” utilized for the control experiment (CTRL).

    b. Microphysics experiments

    The six-species bulk microphysics parameterization described in Lin et al.

    (1983) (hereafter LFO83) was used for all experiments in this study. This scheme

    predicts mixing ratio for rain, hail, snow, graupel, cloud water, and cloud ice. For each

    species, the number concentration is given by

  • 18

    where the subscript x represents the species (e.g., rain, hail, etc.), D is the particle

    diameter, n0 is the intercept parameter, and Λ is the slope parameter. Because the

    LFO83 scheme in ARPS is single-moment, the intercept parameter n0x is specified as a

    constant, and only the slope parameter is computed as a function of state variables.

    Multi-moment schemes in which the intercept parameters vary spatiotemporally have

    shown significant improvements in terms of thermodynamic realism in supercell

    modeling studies (Dawson et al. 2009), but carry too high a computational cost at the

    grid size employed herein to justify their inclusion in this study.

    Among the six species treated in LFO83, parameterizations of rain and hail are

    considered the most important for continental deep moist convection in the

    midlatitudes. Numerous modeling studies, notably Gilmore et al. (2004), van den

    Heever and Cotton (2004), and Snook and Xue (2008) (hereafter SX08), have

    demonstrated a large sensitivity to rain and hail intercept parameters in numerical

    simulations of such convection. Moreover, observational studies such as Waldvogel

    (1974) and Sauvageot and Lacaux (1995) illustrate that these intercept values can vary

    widely in nature, even within a storm. Hence, the choice of n0r and n0h for supercell

    simulations is a nontrivial issue.

    With this in mind, five experiments were performed alongside CTRL whose rain

    and/or hail intercept parameters were varied. These experiments are labeled R86, R85,

    R45, R45H43, and H43, and their respective settings are described in Table 2. The suite

    of Lin ice microphysics experiments in this study follows the general methodology of

    SX08, who also varied the rain and hail intercept parameters in idealized ARPS

    simulations based on the 20 May 1977 sounding and explored the effects on

  • 19

    tornadogenesis. The primary differences here are that we use double the horizontal

    resolution, and the rain and hail intercept parameter values tested in this study cover a

    lower range of values for reasons explained below.

    Note that CTRL does not use the default value for n0r specified in LFO83

    (8x106 m

    -4); instead, it uses a value two orders of magnitude lower than the default

    (8x104 m

    -4). Thus, the labels for the microphysics experiments should be interpreted

    with caution, as CTRL was so labeled simply because its intercept parameter values

    were re-used in all the shear experiments (described in the next subsection), as shown in

    Figure 2. This is because the full set of microphysics experiments was conducted as the

    first stage of this research; afterwards, the experiment with the strongest and most

    persistent low-level mesocyclone was chosen as the control run upon which the shear

    experiments would be based. Comparative plots of domain-wide minimum perturbation

    pressure and maximum low-level vertical vorticity indicated the run with n0r = 8x104

    mm-4

    to be the strongest candidate, resulting in its CTRL designation.

    Experiment R86 uses the default intercept parameter for all species as specified

    in LFO83, and was the first simulation conducted in the present study. In SX08, this is

    regarded as the control experiment, and yields one of the more pronounced tornadic

    vortices within their parameter space. In the present 50 m study, the LFO83-specified

    intercept values in R86 result in very cold and persistent low-level outflow from any

    convective cells developing in the domain. Although focused low-level mesocyclones

    are evident, they are rather brief and contained within air that is quite negatively-

    buoyant. For this reason, the remaining parameter space was chosen to extend only

  • 20

    toward smaller intercept values from R86 with the goal of containing the “ideal” values

    for tornadogenesis within.

    c. Low-level shear experiments

    The literature contains ample evidence linking low-level wind shear with

    tornado potential in supercell environments. In the context of supercells, the low-level

    shear is typically characterized by one of two means which are related but distinct. The

    first uses the perspective of storm-relative helicity (SRH), which quantifies

    environmental streamwise vorticity over a specified depth of the atmosphere. SRH,

    which is a function of both the environmental shear and storm motion, has been a

    popular tool to discriminate between tornadic and nontornadic storms over the past few

    decades, and as such many studies have explored its utility in this regard, e.g.

    Markowski et al. (1998). One practical issue with using SRH diagnostically, particularly

    before a storm has even formed, is its strong sensitivity to storm motion, which may

    vary widely in time and space (Weisman and Rotunno 2000). A second, and perhaps

    currently less-popular, means to characterize low-level shear is simply its magnitude as

    a vector difference per unit depth. The shear value itself does not incorporate the effects

    of streamwise vorticity, which can be both an advantage (less spatiotemporally variable

    and complex to calculate) and disadvantage (may fail to identify tornado likelihood

    when or if environmental vorticity is an important source). Though the body of work

    relating this quantity to tornadogenesis is comparatively smaller than that for SRH, a

    few pertinent studies are addressed below.

    Using a relatively low-resolution numerical simulation, Weisman and Klemp

    (1982) (hereafter WK82) explored the relationship between vertical shear magnitude

  • 21

    and storm behavior for unidirectional westerly flow, finding a distinct maximum in low-

    level vertical vorticity production at a 0-10 km vector wind difference of 25 m s-1

    . Of

    significance is that even higher shear values (35 m s-1

    and 45 m s-1

    ) yielded drastically

    less intense cyclonic vertical vorticity near the ground, despite stronger midlevel

    cyclonic vorticity. Note that in the experiments of WK82, the wind speed was

    proportional to the hyperbolic tangent of height AGL, placing most of the shear in the

    lowest few kilometers. As such, their results can be interpreted primarily as describing a

    “sweet spot” at which low-level wind shear maximizes low-level vorticity within

    supercells.

    Markowski et al. (2003) examined an extensive dataset of RUC model proximity

    soundings for strongly-tornadic, weakly-tornadic and nontornadic supercells. They

    explored both SRH and shear magnitude in relation to these categories, finding both to

    be convincingly correlated with tornado likelihood and strength in the 0-1 km AGL

    layer. They do note that “[n]o robust differences in hodograph curvature exist” among

    these three categories, implying that the bulk shear’s contribution to SRH could

    potentially be more important than the precise orientation of the shear. In the same vein,

    Esterheld and Giuliano (2008) examined proximity soundings which revealed that in the

    10-1000 m AGL layer, shear magnitude better discriminated between strongly- and

    weakly-tornadic supercells than SRH.

    Schumacher and Boustead (2011) (hereafter SB11) examined the 24 June 2003

    tornado outbreak in Nebraska and South Dakota, and found in particular that an increase

    in the magnitude of the 0-1 km AGL shear vector played a crucial role in the transition

    from nontornadic or weakly tornadic supercells early in the event to strongly tornadic

  • 22

    supercells later. They emphasize that even when total shear through the storm-bearing

    layer (0-6 km AGL) remained steady or decreased slightly, a redistribution of vertical

    wind shear into the lowest kilometer proved quite favorable for significant tornadoes.

    In order to shed further light on the role of low-level shear on supercell tornado

    formation, five experiments were performed alongside CTRL whose initial

    environments were characterized by kinematic modifications to the “may20” sounding

    in the lowest 2 km AGL. These experiments (and their respective soundings) are labeled

    IS1 (may20_lls2), IS2 (may20_lls3), IS3 (may20_lls4), DS1 (may20_lls5), and DS2

    (may20_lls6). Figure 3 and Figure 4 display vertical profiles below 2 km AGL for the

    u- and v-components of the horizontal wind, respectively, for these soundings. For

    numerical values, reference Appendix A. Comparable vertical profiles for the supercell

    cases of Markowski et al. (2003) discussed earlier in this subsection are presented in

    Figure 5, though the plots extend upward to 12 km AGL and values are ground-relative

    (rather than storm-relative).

    The kinematic differences amongst these initial soundings are characterized by

    varying vertical distributions of wind shear in the lowest 2 km of the model domain,

    while preserving approximately the same hodograph shape. The surface wind vector is

    held constant for all experiments. Experiments IS1, IS2 and IS3 feature additional wind

    shear in the 0-1 km AGL layer, with a compensating decrease in the 1-2 km AGL layer.

    Conversely, experiments DS1 and DS2 feature weaker wind shear in the 0-1 km AGL

    layer, with a compensating increase in the 1-2 km AGL layer. All wind values are

    identical between experiments at and above 1750 m AGL (note that some are identical

    to one another at levels lower than 1750 m AGL; consult Figure 3 and Figure 4 for

  • 23

    details). To facilitate comparisons with the aforementioned work in the literature,

    vertical shear magnitude in the 0-500 m AGL and 0-1000 m AGL layers was calculated

    for each experiment and presented in Table 3.

    Experiment Sounding n0r (m-4

    ) n0h (m-4

    )

    CTRL may20 8x104

    4x104

    R86 may20 8x106

    4x104

    R45 may20 4x105

    4x104

    R85 may20 8x105 4x10

    4

    R45H43 may20 4x105

    4x103

    H43 may20 8x104

    4x103

    IS1 may20_lls2 8x104

    4x104

    IS2 may20_lls3 8x104

    4x104

    IS3 may20_lls4 8x104

    4x104

    DS1 may20_lls5 8x104

    4x104

    DS2 may20_lls6 8x104

    4x104

    Table 2. Summary of soundings and intercept values for all experiments.

    Figure 2. Graphical depiction of parameter space covered by all experiments.

  • 24

    Experiment 0-500 m AGL shear (s-1

    ) 0-1000 m AGL shear (s-1

    )

    CTRL 0.0026 0.0067

    IS1 0.0032 0.0067

    IS2 0.0040 0.0070

    IS3 0.0057 0.0075

    DS1 0.0018 0.0055

    DS2 0.0010 0.0043

    Table 3. Low-level vertical wind shear magnitudes for all shear experiments.

    Figure 3. Zonal (u) wind component for the six initial soundings.

  • 25

    Figure 4. Meridional (v) wind component for the six initial soundings.

    Figure 5. Average zonal (u) and meridional (v) wind components for strongly-tornadic

    (solid), weakly-tornadic (dashed), and nontornadic (dotted) supercell environments as

    derived from 413 RUC proximity soundings. Adapted from Markowski et al. (2003).

  • 26

    3. Definitions

    a. Mesocyclone

    Wicker and Wilhelmson (1995) (hereafter WW95) define a mesocyclone, at any

    horizontal slice through a supercell thunderstorm, as the region within which vertical

    vorticity exceeds 0.01 s-1

    . Owing to the high horizontal resolution of simulations in our

    study relative to earlier modeling-based work, however, this definition is somewhat

    problematic in application. Through the life cycles of our modeled supercells, the

    presence of noise and other small-scale features on the order of 100 m (2Δx) often

    precludes the identification of contiguous regions inside the ζ = 0.01 s-1

    contour which

    are of a radius consistent with observed low-level mesocyclones (i.e., a few kilometers).

    Even where the flow pattern features a circulation suggestive of a mesocyclone, the

    embedded vertical vorticity distribution typically contains small patches of low (and

    even negative) values. For this reason, the identification of mesocyclones hereafter is

    somewhat qualitative, while attempting to conform to the spirit of the vorticity-based

    definition.

    b. Tornado

    Identification of tornadoes within these simulations presents similar challenges

    to those regarding mesocyclones. A wind speed threshold alone is of limited utility

    because it fails to account for the presence of strong circulation and vorticity. A

    vorticity threshold, while more pertinent, is impractical due to the high dependence of

    derivative quantities (e.g., du/dx) upon grid spacing. Pressure deficit is a more attractive

    and practical choice for setting a threshold, as it is predicted explicitly by the model and

  • 27

    not as sensitive to grid spacing. WW95 demonstrate that for a Rankine combined vortex

    in cyclostrophic balance,

    where Δp is the central pressure deficit, ρ is the air density, and Vt is the maximum

    tangential wind speed. Clearly, the vortices which develop in our simulations will

    neither be in perfect cyclostrophic balance nor contain uniform tangential winds at a

    given radius. Nevertheless, this approximation is useful for choosing a threshold which

    introduces an objective element into the tornado-identification process. WW95 set a

    minimum surface wind speed threshold of 30 m s-1

    ; considering this value for Vt in the

    equation and assuming ρ ~ 1 kg m-3

    yields a pressure deficit of 9 hPa. Taking this into

    consideration, the following criteria are used in order to classify a vortex as tornadic:

    1) The maximum surface pressure deficit is at least 10 hPa.

    2) The maximum embedded surface wind speed is at least 30 m s-1.

    3) The radius of the vortex is less than 750 m.

    4) The surface flow pattern is approximately circular.

    Criterion 3 is supported by Bluestein (2007), an extensive literature review

    which found a generally-accepted maximum tornado diameter of ~1.5-2 km. It must be

    clarified that Criterion 4 references only the wind vector directions; there is no

    expectation that the distribution of wind speeds will be axisymmetric. Criteria 3 and 4

    introduce subjectivity due to the nebulous definitions of “radius” and “circular,”

    respectively, but this is necessary in order to avoid classifying diffuse and/or

    disorganized low-level mesocyclones (which might contain transient vortices with

    substantial pressure drops) as tornadoes.

  • 28

    Chapter 4: Analysis of Control Simulation

    1. Storm morphology

    The initial thermal bubble initiates deep moist convection rapidly in CTRL, and

    the domain-wide maximum reflectivity exceeds 30 dBZ by 600 s. This cell initially

    translates rapidly to the northwest, influenced primarily by flow in the lowest few

    kilometers owing to its shallow vertical extent. A convective storm with ground-level

    reflectivity exceeding 60 dBZ is evident by 1800 s near the center of the domain (x = 32

    km, y = 32 km).

    As early as 3000 s, the initially-unicellular updraft appears to have split in the

    manner described by Klemp and Wilhelmson (1978), with one cell to the west and

    another to the east (Figure 6a,Figure 7a). While the eastern cell slows significantly

    during the period 3000-4200 s, the western cell continues translating rapidly to the

    west-northwest, eventually reaching the western boundary of the domain. The western

    cell during this period is considered to be the left-moving split of the initial convective

    storm, while the eastern cell is the right-moving split. The left split later weakens and

    ultimately dissipates at the northwest corner of the model domain by 8000 s,

    presumably owing in large part to its advection out of the domain. The right split is the

    focus of the analysis hereafter, and will be labeled “Cell A.”

    Cell A propagates westward to a position near (x = 24 km, y = 36 km) by 3600

    s. At this time, supercellular characteristics are already evident, including a well-defined

    FFD and RFD in the ground-level reflectivity field. The horizontal extent of the storm,

    as defined by the 30 dBZ reflectivity contour at ground level, is approximately 30x16

    km. A surge of relatively cold outflow at ground level emanates from the storm’s FFD

  • 29

    around 4500 s, and this outflow expands southward and westward away from the storm

    over the next several minutes.

    During the period 4800-5400 s, a new convective cell initiates near the

    southwest corner of the domain, which will hereafter be called “Cell B.” Based upon a

    movie of ground-level potential temperature and wind vectors (not shown), it appears

    that outflow from early in Cell A’s life cycle was the source of low-level convergence

    that instigated this development. Between 5400 s and 8000 s, Cell B remains nearly

    stationary, with its precipitation core at ground-level centered near (x = 12 km, y = 20

    km). Meanwhile, Cell A propagates south-southwestward, and its own precipitation

    core eventually begins to merge with Storm B around 8400 s. Their respective mid-level

    updrafts also merge around 9000-9600 s, coinciding with significant intensification of

    the low-level mesocyclone originally associated with Cell A. This raises the possibility

    that the storm merger played a role in tornadogenesis, perhaps by modulating the

    strength and storm-relative position of the primary updraft. This issue is further

    explored in the next subsection.

    Prior to the storm merger, the Cell A exhibits well-defined supercell structure,

    including a hook echo in the ground-level reflectivity field around 4800-5400 s (Figure

    6b). Despite this precipitation configuration, the surface flow pattern is not indicative of

    a well-formed ground-level mesocyclone; accordingly, positive surface vertical

    vorticity is not present in noteworthy quantity or concentration at the hook. Above the

    surface, the flow pattern at 1039 m AGL does suggest the presence of a mesocyclone,

    and a concentrated region of vertical vorticity exceeding 0.01 s-1

    exists at the tip of the

  • 30

    hook. Therefore, this first hook echo is associated with a low-level mesocyclone, but

    one which does not reach the ground.

    Figure 6. Perturbation potential temperature (shading; K) and 1 g kg-1

    rainwater mixing

    ratio contour at 10 m AGL for (a) 3000 s, (b) 4800 s, (c) 7200 s, and (d) 10200 s.

  • 31

    Figure 7. As in Fig. 5, but for vertical velocity (shading; m s-1

    ) at 4265 m AGL.

    The initial hook echo becomes ill-defined by 6000 s as precipitation in the RFD

    of Cell A expands in horizontal extent. During this period leading up to the storm

    merger, positive values of low-level vertical vorticity are found in a broad zone along

    the primary storm’s forward- flank and rear-flank gust front (FFGF and RFGF,

    respectively). This period also marks a transition in the storm’s spatial configuration.

    Initially, the RFD and FFD storm-relative locations approximately followed the

    conceptual model of Lemon and Doswell (1979): to the west and northeast of the main

    updraft, respectively. By 7200 s, however, the entire supercell appears rotated 45 to 90

    degrees counterclockwise from the classic conceptual model: specifically, the FFGF

  • 32

    extends northward from its intersection with the RFGF, rather than northeastward or

    eastward (Figure 6c). Indeed, the highest reflectivity marking the FFD core is located

    northwest of the gust front intersection, rather than northeast. The strongly-meridional

    environmental mid-level wind on the initial sounding likely helps to account for this, as

    the entire (ground-relative) wind profile is somewhat more backed than the classic

    southwest flow environment upon which conceptual models were built. This rotated

    configuration persists for the remainder of the simulation.

    Immediately after the storm merger and associated low-level mesocyclone

    intensification of the 9300-9600 s period, a surge of cold air stronger than previous

    episodes (surface potential temperature perturbations on the order of -12 to -15 K)

    emanates from the RFD at ground-level and spreads quickly in the horizontal. This

    rapidly occludes a much longer segment of the gust front, with deep-layer downdraft

    concurrently overspreading the mesocyclone. By 10200 s, cold air at the surface covers

    much of the southwest quadrant of the domain behind the RFGF, which surges well

    southeast of the low-level mesocyclone (Figure 6d). Simultaneously, the mid-level

    updraft appears less cellular, instead becoming elongated and bow-shaped

    approximately above the length of the RFGF (Figure 7d). For the remainder of the

    simulation, the mid-level updraft is relatively weak and cold air at the surface extends to

    the southern boundary of the domain south of the storm, indicating it is no longer

    containing its own outflow.

  • 33

    2. Low-level mesocyclone intensification and tornadogenesis

    a. Evolution of model fields

    In CTRL, a mid-level mesocyclone is detected as early as 1800 s, even before

    the initial convective cell has split. Given effective-layer storm-relative helicity (SRH)

    exceeding 100 m2 s

    -2, this is expected, as updraft is correlated with positive vertical

    vorticity in such an environment. As described by Davies-Jones et al. (2001), only

    afterwards does a low-level mesocyclone develop. The first evidence of a low-level

    mesocyclone appears around 4800 s, when a region of positive vertical vorticity at 1039

    m AGL becomes concentrated near the tip of Cell A’s hook echo. This region of

    vorticity persists in some form for the remainder of the period leading up to the storm

    merger, but at times grows more diffuse and elongated along the gust front.

    As the storm merger begins in earnest around 9000 s, numerous compact lobes

    of > 0.1 s-1

    vertical vorticity are present along Cell A’s FFGF and RFGF at the surface,

    with several in close horizontal proximity to the gust front intersection (Figure 8a).

    Over the 300 s that follow, lobes of vorticity continue to advect southward down the

    FFGF and concentrate vorticity at the intersection with the RFGF; simultaneously, a

    surge of relatively strong northwesterly flow develops in the RFD just to the northwest

    of the gust front intersection. Plots of perturbation potential temperature reveal this

    surge to be associated with air that is slightly warmer than the surrounding FFD and

    RFD. Between 9300-9500 s, this surge begins to wrap cyclonically around the south

    and east sides of the low-level mesocyclone. This forces the northernmost segment of

    the RFGF to pivot northwestward, resulting in an occlusion on the north side of the

    mesocyclone (Figure 8b).

  • 34

    Figure 8. Close-up (moving) view of low-level mesocyclone at 10 m AGL.

    Perturbation potential temperature (shading; K) and 0.1 s-1

    vertical vorticity contour

    (orange solid) are plotted at (a) 9000 s, (b) 9420 s, (c) 9600 s, and (d) 9900 s. Gust

    fronts are indicated by solid red (FFGF), dark blue (RFGF), and purple (occluded)

    curves. Black “T” indicates tornado position.

    Just to the southeast of the occlusion point, a tornado develops around 9480 s. A

    24x24 km subdomain centered approximately on the tornado track was extracted from

    the 64x64 km model domain, and all spatial references for the remainder of this section

    pertain to this subdomain, denoted by xs and ys. (The subdomain is not a nested grid nor

    does it contain finer-scale interpolation, and was produced simply to reduce the

    computational demands of reading data for plots and trajectory calculations). The

    tornado develops near (xs = 11.0 km, ys = 13.6 km) in the subdomain. This vortex

  • 35

    results from the merger of two pretornadic vortices which can be identified as early as

    9180 s, or three minutes before tornadogenesis (Figure 9a). At this stage, low-level

    updraft is present above the southern portion of the stronger pretornadic vortex (labeled

    “PV1”), but the strongest updraft remains on the west side of the mesocyclone (Figure

    9e). By 9480 s, the low-level updraft has become more cellular and the strongest portion

    has advanced cyclonically to a position over the southern portion of the incipient

    tornado (Figure 9f). The tornado initially has a horizontal radius of approximately 200

    m and tightly-concentrated pressure drop at the surface at 9480 s (Figure 9b), but with

    time evolves into a broader circulation with a radius around 400 m by 9600 s (Figure

    9c). This evolution coincides with the RFGF continuing to surge outward from the low-

    level mesocyclone, forcing its occlusion point with the FFGF farther to the north

    (Figure 8c). By 9720 s, despite maintaining a surface pressure deficit exceeding 10 hPa,

    the vortex is no longer considered tornadic due to its radius appearing to exceed 750 m

    (Figure 9d). Low-level updraft over the southern half of the vortex has eroded, while

    downdraft has expanded and strengthened over its northern half (Figure 9h). While

    surface wind speeds at that time exceed 35 m s-1

    at the radius of maximum winds, a

    large circular zone (radius around 500 m) of winds less than 15 m s-1

    exists at the vortex

    center. Between 9600-9700 s, immediately prior to this expansion of the vortex into a

    more diffuse and disorganized entity, a second, colder RFD surge appears to wrap

    cyclonically around and into the surface circulation.

  • 36

    Figure 9. Zoomed (moving) view of the tornado in CTRL. (a-d) Horizontal wind speed

    (shading; m s-1

    ) and perturbation pressure at 3 hPa increments below -3 hPa (contour),

    at 10 m AGL. (e-h) Vertical velocity (shading; m s-1

    ) and p’ (contour; -5 hPa) at 1039 m

    AGL. Pretornadic vortices are denoted by “PV1” and “PV2” in (a).

  • 37

    To compliment the horizontal plots, vertical cross-sections in the xz-plane are

    presented in Figure 10. Note that the position and orientation of these slices are chosen

    manually at each time to follow the vortex tilt with height. At 9180 s, three minutes

    before the time at which we declare tornadogenesis to occur, the stronger of the two

    pretornadic vortices appears formidable in its own right. A vertical slice from (xs = 8, ys

    = 15) to (xs = 14, ys = 14.6) reveals the vortex (as defined by the 0.1 s-1

    vertical

    vorticity contour) to extend from the surface up to around 3 km AGL, with a 20-30° tilt

    toward the west with height (Figure 10a). While vertical velocity inside the vortex is

    small in magnitude, strong updraft exceeding 35 m s-1

    is found in the mid-levels

    immediately above. By 9360 s, however, an equivalent cross-section following the

    vortex (not shown) shows that this vortex has weakened in the 1-3 km AGL layer; the

    0.1 s-1

    vertical vorticity contour only extends up to about 900 m AGL, though mid-level

    updraft immediately above remains strong. Entering the tornadic stage at 9480 s, the

    vortex regains a depth of nearly 3 km, while mid-level updraft immediately above has

    weakened slightly (Figure 10b). Given the somewhat broken, patchy distribution of

    vertical vorticity, it might be inferred that the tornado is comprised of multiple vortices.

    However, close inspection of the horizontal wind field at grid scale confirms that the

    circulation, while broad, is a single, continuous entity, with no embedded sub-vortices

    apparent. Local vorticity maxima within owe largely to small shear zones, which could

    hint at subvortices the 50 m horizontal grid spacing is unable to resolve explicitly.

    Significant changes appear by 9600 s, when the tornado is best-organized and

    exhibits the strongest surface pressure deficit. A vertical slice through from (xs = 9, ys =

    10.3) to (xs = 15, ys = 15) shows the tornado contains deep, strong downdraft driven by

  • 38

    its own downward-directed perturbation pressure gradient force, signaling its imminent

    demise (Figure 10c). Indeed, by 9720 s, a vertical slice reveals this downdraft has

    further strengthened and that the original vortex has weakened substantially near the

    surface (Figure 10d). A strong low-level mesocyclone exists through the 1-3 km AGL

    layer, but it contains significant downdraft which is likely contributing to the

    divergence and less-organized structure at the surface.

    Some degree of central downdraft is observed throughout the tornado’s life

    cycle, and there exists a core of relatively weak tangential flow which expands radially

    outward with time until the vortex can no longer be considered tornadic. It is speculated

    that the lack of surface friction in the simulation may be a primary driver of this

    structure. Trapp (2000) observed similar structure when simulating a vortex using a

    free-slip lower boundary condition, labeling it a “two-celled vortex." In such a vortex,

    friction is not allowed to disrupt cyclostrophic balance near the surface, and low-level

    updraft is confined to just outside the vortex core. A schematic is presented in Figure

    11, revealing a vertical velocity distribution analogous to that seen in Figure 10 (in

    particular, Figure 10c). Thus, the vortex in CTRL from 9480-9720 s is considered a

    two-celled vortex. Observations indicate this configuration is typically seen late in the

    life cycles of some tornadoes, after vortex breakdown has occurred (Trapp 2000). In

    CTRL, however, no such breakdown occurs. Observations of two-celled tornadoes in

    nature do exist (Lee and Wurman 2005), but appear somewhat rare. An issue to explore

    in future work is why such structure is seen here, while some other free-slip simulations

    using the 20 May 1977 sounding (e.g., NN10) produce central updraft until late in the

    tornado life cycle.

  • 39

    Figure 10. Vertical cross-section through (a) pretornadic vortex PV1 at 9180 s, and (b-

    d) the tornado at 9480, 9600, and 9720 s. Vertical velocity (shading; m s-1

    ) is shown,

    along with positive vertical vorticity contours (s-1

    ) at 0.1 s-1

    intervals starting at 0.1 s-1

    .

    The horizontal endpoints of each slice are indicated as (xs, ys).

    Figure 11. Conceptual model of a two-cell vortex associated with subcritical flow.

    Adapted from Trapp (2000).

  • 40

    Time-height cross sections are presented for subdomain-wide maximum vertical

    vorticity (Figure 12a) and maximum updraft (Figure 12b). In light of the analyses

    performed via instantaneous cross sections above, it is immediately apparent that

    caution must be exercised when interpreting these time-height plots for a couple

    reasons. First, the presence of two distinct supercells (Cell A and Cell B) in close

    proximity complicates interpretation of domain-wide maximum and minimum values,

    potentially masking important trends in one storm or the other. (Furthermore, choosing

    separate subdomains for each cell over which to compute extrema is impractical, since

    they merge with time). This may explain why any individual updraft pulses and

    vorticity maxima prior to the merger at 9000 s appear fairly nondescript in the extrema

    plots. Second, the strongest low-level vertical vorticity in the simulation occurs from

    9960-10200 s, but horizontal cross-sections reveal this maximum owes to transient,

    shallow vortices along the FFGF; these might be considered strong gustnadoes. The

    more sustained tornado we follow between 9480-9720 s is coincident with another,

    weaker maximum in vertical vorticity below 1 km AGL, as well as a period of relatively

    strong low-level updraft exceeding 30 m s-1

    . One notable difference between Figure 12

    and comparable plots in NN10 and others is the relatively abrupt appearance of the

    primary pressure drop over a deep layer around 9100 s. Whereas earlier studies

    examining prototypical supercell evolution found a mid-level mesolow which expanded

    vertically with time (over tens of minutes) to support tornadogenesis, the low-level

    updraft and deep-layer perturbation pressure minimum here appear quite suddenly, only

    a few minutes before tornadogenesis. It appears the merger of updraft associated with

    Cell A and Cell B around this time aided in this abrupt transition.

  • 41

    Figure 12. Time-height cross-section of (a) maximum vertical velocity, (b) maximum

    vertical vorticity, and (c) minimum perturbation pressure over the model domain from

    7200-12600 s.

  • 42

    The precise means by which the storm merger enhanced low-level

    mesocyclogenesis and ultimately led to tornadogenesis will require further work in the

    future, although both time-height and instantaneous horizontal cross sections clearly

    reveal low-level updraft intensification coincident with the merger. Lee et al. (2006)

    documented a significant temporal association between storm mergers and tornado

    incidence during an Illinois outbreak in 1996, speculating that perhaps outflow

    boundaries separating air with relatively small differences in θe (which were observed in

    their case) could provide constructive interference to a mature supercell. In this study,

    however, the younger storm (Cell B) does not appear to feature a well-formed, distinct

    gust front antecedent to the merger with which the more mature storm (Cell A) could

    interact. Wurman et al. (2007) also observed repeated tornadogenesis during mergers in

    a 1997 supercell episode in Oklahoma, hypothesizing a corresponding increase in low-

    level convergence and updraft to be the causal factor; the results modeled in this study

    would support this idea.

    b. Trajectory analysis

    To investigate the spatial origins of air parcels entering the low-level

    mesocyclone and tornado, backward time-dependent parcel trajectories were calculated.

    The model was re-run to obtain 3 s data within the time period of interest (9000-10200

    s), and several initial times were selected from which to integrate backward trajectories.

    For each initial time chosen, 20 trajectories were initialized along a ring centered within

    the low-level mesocyclone (or tornado, if it was present at the initial time), along with

    one trajectory initialized at the center of that ring. The ring radius was 200 m, and initial

    parcel heights were 50 m. Backward calculations were carried out in time to 480 s prior

  • 43

    to initialization using Heun’s method and trilinear spatial interpolation between grid

    points; see Appendix B for a complete description of the trajectory calculation

    procedure. These calculations utilized wind data dumped every 3 s from the model with

    no temporal interpolation. Some implications of this choice are discussed in Chapter 7.

    Figure 13a illustrates 21 representative parcels entering the tornado at 9480 s,

    near the time of tornadogenesis. All but one of these parcels originates from the

    northwest of the tornado in the FFD region; the lone exception appears to be a parcel

    which followed the stronger pretornadic vortex for several minutes prior. At 8 min prior

    to initialization, there is a substantial range of original parcel heights in the FFD.

    Several parcels originate from above 1000 m AGL, while others move primarily in the

    horizontal near the ground. Most parcels in the latter group exhibit spiraling paths in the

    horizontal as they approach the tornado which suggest they may have entered the

    pretornadic vortex a minute or two prior; by contrast, the only parcels with a relatively

    straight path into the tornado at 9480 s originated from above 1000 m AGL. Thus, it is

    hypothesized that the wide disparity in original heights owes significantly to a

    corresponding disparity in the time at which parcels entered the vortex, and that

    ultimately most parcels participating in the tornado descended in the FFD at some time

    prior. The typical parcel path can then be described as first descending in the FFD, then

    moving southeast below 100 m AGL through the RFD region and ultimately into the

    tornado from the south and west.

  • 44

    Figure 13. Parcel trajectories initialized in the vortex at (a) 9480 s, (b) 9600 s, and (c)

    9780 s; and in the low-level mesocyclone center at (d) 10080 s. Parcels were initialized

    at 50 m AGL and integrated backward in time for 480 s. Parcel height is indicated by

    color, providing a pseudo-3D visualization.

    Figure 13b shows parcels within the tornado at 9600 s, around the time of its

    peak strength. A few notable changes are evident. First, a relatively high proportion of

    the parcels appear to have followed the tornado over the preceding minutes, rather than

    entering it at this time. Additionally, there are now two parcels with “inflow

    trajectories” from the east of the vortex, where none existed at 9480 s. In an intensive

    study of trajectory calculation methodology in the context of supercell mesocyclones

    and tornadoes, Dahl et al. (2012) found that such trajectories often result from artifacts

    of the backward-integration methodology, and are not present when more-accurate,

  • 45

    equivalent forward trajectories are calculated using the model time step. For this reason,

    the validity of these parcel trajectories is called into question and they will be

    discounted. Among the remaining parcels, the group originating above 500 m AGL

    appears now to descend from west of the vortex in the RFD, rather than northwest in the

    FFD, as was seen at 9480 s. This trend continues at 9780 s (Figure 13c), at which time

    the vortex has become diffuse and nontornadic. Thus, over the lifespan of the tornado,

    the area of origin for descending participating parcels moves cyclonically around the

    low-level mesocyclone, from northwest to west-southwest of the tornado. As illustrated

    in Figure 13d, parcels entering the broad, divergent low-level mesocyclone at 10080 s

    descend sharply from nearly overhead, as strong downdraft has now overtaken the

    entire low-level mesocyclone.

    To clarify the reason for the cyclonic wrapping of the parcel-origin region, plots

    of vertical pressure gradient force (not shown) were inspected, and did not reveal

    significant dynamic forcing for descent in the parcel-origin areas. However, a maximum

    in the reflectivity field at 1 km AGL was observed to wrap cyclonically around the low-

    level updraft between 9000-9600 s, coinciding with the evolution of parcel origin

    (Figure 14). Horizonta


Recommended