Skip to main content

Advertisement

Log in

Tracking the Centre of Asymmetric Vortices Using Wind Velocity Vector Data Fields

  • Research Article
  • Published:
Boundary-Layer Meteorology Aims and scope Submit manuscript

Abstract

Tornados are a major hazard in many regions around the world and as such it is necessary to analyze them. However, such analyses require accurately tracking them first. Currently, there are gaps in the available vortex detection methods when processing a wind-field dataset to locate a series of points that are identifiable as the tornado centreline. This study proposes a novel solution that corrects for deficiencies in previous attempts to identify vortex centres when applied to tornado wind-fields, which would have otherwise led to identifying merely the region of the vortex, several potential centres requiring post-processing, or erroneously approximating the tornado centre. Additionally, this method combines the efficiency required to process large datasets of temporal and spatial wind velocity vector distributions with the accuracy needed to reliably calculate a specific line as a tornado centre. This method is compared to five other approaches commonly used for vortex identification in order to assess: (a) how accurately they identify the centre region, (b) how they handle extraneous vortices that are not of interest, and (c) their computational efficiency in processing a wind-field dataset. With the proposed method, it would be possible to plot a tornado path from formation to dissipation and perform analyses to understand the vortex characteristics with respect to this path without requiring extensive user-intervention.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  • Aboelkassem Y, Vatistas GH, Esmail N (2005) Viscous dissipation of Rankine vortex profile in zero meridional flow. Acta Mech Sinica 21(6):550–556

    Article  Google Scholar 

  • Agee E, Taylor L (2019) Historical analysis of US tornado fatalities (1808–2017): population, science, and technology. Weather Clim Soc 11(2):355–368

    Article  Google Scholar 

  • Altalmas A, El Damatty AA (2014) Finite element modelling of self-supported transmission lines under tornado loading. Wind Str 18(5):473–495

    Article  Google Scholar 

  • Baker CJ, Sterling M (2017) Modelling wind fields and debris flight in tornadoes. J Wind Eng Ind Aerodyn 168:312–321

    Article  Google Scholar 

  • Beck V, Dotzek N (2010) Reconstruction of near-surface tornado wind fields from forest damage. J Appl Meteorol Climatol 49(7):1517–1537

    Article  Google Scholar 

  • Brooks HE, Correia J (2018) Long-term performance metrics for national weather service tornado warnings. Weather Forecast 33(6):1501–1511

    Article  Google Scholar 

  • Bryan G (2008) CM1 Homepage. In: University Corporation for Atmospheric Research. http://www2.mmm.ucar.edu/people/bryan/cm1/. Accessed 29 June 2020

  • Bryan GH, Fritsch JM (2002) A benchmark simulation for moist nonhydrostatic numerical models. Mon Weather Rev 130(12):2917–2928

    Article  Google Scholar 

  • Chen Q, Zhong Q, Qi M, Wang X (2015) Comparison of vortex identification criteria for planar velocity fields. Phys Fluids 27(8):085101

    Article  Google Scholar 

  • Church CR, Snow JT, Baker GL, Agee EM (1979) Characteristics of tornado-like vortices as a function of swirl ratio: a laboratory investigation. J Am Meteorol Soc 36(9):1755–1776

    Google Scholar 

  • Curtis A, Mills JW (2012) Spatial video data collection in a post-disaster landscape: the Tuscaloosa tornado of April 27th 2011. Appl Geogr 32(2):393–400

    Article  Google Scholar 

  • Davies-Jones R (2008) Can a descending rain curtain in a supercell instigate tornadogenesis barotropically? J Atmos Sci 65(8):2469–2497

    Article  Google Scholar 

  • Doswell CA, Moller AR, Brooks HE (1999) Storm spotting and public awareness since the first tornado forecasts of 1948. Weather Forecast 14(4):544–557

    Article  Google Scholar 

  • Finley CA, Orf L, Lee BD, Wilhelmson RB (2018) High-resolution simulation of a violent tornado in the 27 April 2011 outbreak environment. In: Proceedings of the 29th Conference on Severe Local Storms, Pinnacle AB. Am Meteorol Soc. https://ams.confex.com/ams/29SLS/webprogram/Paper348812.html. Accessed 1 Sep 2020

  • Grams JS, Thompson RL, Snively DV, Prentice JA, Hodges GM, Reames LJ (2012) A climatology and comparison of parameters for significant tornado events in the United States. Weather Forecast 27:106–123

    Article  Google Scholar 

  • Hamada A, El Damatty AA (2011) Behaviour of guyed transmission line structures under tornado wind loading. Comput Str 89(11–12):986–1003

    Article  Google Scholar 

  • Jeong J, Hussain F (1995) On the identification of a vortex. J Fluid Mech 285:69–94

    Article  Google Scholar 

  • Jiang M, Machiraju R, Thompson D (2002) A novel approach to vortex core region detection. In: Elbert D, Brunet P, Navazo I (eds) Joint Eurographics-IEEE TCVG symposium on visualization, vol. 2, pp 217–225

  • Jiang M, Machiraju R, Thompson D (2005) Detection and visualization of vortices. In: Hansen CD, Johnson CR (eds) The visualization handbook. Elsevier Academic Press, Oxford, pp 295–309

    Chapter  Google Scholar 

  • Karstens CD, Gallus WA, Lee BD, Finley CA (2013) Analysis of tornado-induced tree fall using aerial photography from the Joplin, Missouri, and Tuscaloosa-Birmingham, Alabama, tornadoes of 2011. J Appl Meteorol Climatol 52(5):1049–1068

    Article  Google Scholar 

  • Kuo HL (1971) Axisymmetric flows in the boundary layer of a maintained vortex. J Atmos Sci 28:20–41

    Article  Google Scholar 

  • Levy Y, Degani D, Seginer A (1990) Graphical visualization of vortical flows by means of helicity. AIAA J 28(8):1347–1352

    Article  Google Scholar 

  • Lewellen WS (1993) Tornado vortex theory. In: Church C, Burgess D, Doswell C, Davies-Jones R (eds) The tornado: its structure, dynamics, predictions, and hazards. Am Geophys Union, Washington, 79:19–39

  • Lim JR, Liu BF, Egnoto M (2019) Cry wolf effect? Evaluating the impact of false alarms on public responses to tornado alerts in the southeastern United States. Weather, Clim, Soc 11(3):549–563

    Article  Google Scholar 

  • Liu Z, Zhang C, Ishihara T (2018) Numerical study of the wind loads on a cooling tower by a stationary tornado-like vortex through LES. J Fluids Str 81:656–672

    Article  Google Scholar 

  • Natarajan D (2011) Numerical simulation of tornado-like vortices. PhD dissertation, The University of Western Ontario, Ontario, Canada

  • National Centers for Environmental Information [NCEI] (2021a) Billion-dollar weather and climate disasters. Natl Oceanic Atmos Adm. https://www.ncdc.noaa.gov/stormevents/. Accessed 2 February 2022.

  • National Centers for Environmental Information [NCEI] (2021b) Storm events database. Natl Oceanic Atmos Adm. https://www.ncdc.noaa.gov/billions/events/US/1980-2021b. Accessed 2 February 2022.

  • Orf L (2019) A violently tornadic supercell thunderstorm simulation spanning a quarter-trillion grid volumes: computational challenges, I/O framework, and visualizations of tornadogenesis. Atmos 10(10):578–601

    Article  Google Scholar 

  • Orf LG, Semeraro BD, Wilhelmson RB (2007) Vortex detection in a simulated supercell thunderstorm. Atmos Sci Lett 8:29–35

    Article  Google Scholar 

  • Orf L, Wilhelmson R, Wicker L (2016) Visualization of a simulated long-track EF5 tornado embedded within a supercell thunderstorm. Parallel Comput 55:28–34

    Article  Google Scholar 

  • Pecin TG, Almeida AAD, Roehl JL (2011) Tornadic mechanical global actions on transmission towers. J Brazilian Soc of Mech Sci Eng 33(2):131–138

    Article  Google Scholar 

  • Potvin CK (2013) A variational method for detecting and characterizing convective vortices in Cartesian wind fields. Mon Weather Rev 141(9):3102–3115

    Article  Google Scholar 

  • Rankine WJM (1877) Motions of fluids – hydrodynamics. Manual of Appl Mech. Charles Griffin and Company, London, pp 574–577

    Google Scholar 

  • Razavi A, Sarkar PP (2018) Laboratory study of topographic effects on the near-surface tornado flow field. Boundary-Layer Meteorol 168(2):189–212

    Article  Google Scholar 

  • Refan M (2014) Physical simulation of tornado-like vortices. PhD dissertation, The University of Western Ontario, Ontario, Canada

  • Refan M, Hangan H (2016) Characterization of tornado-like flow fields in a new model scale wind testing chamber. J Wind Eng Ind Aerodyn 151:107–121

    Article  Google Scholar 

  • Refan M, Hangan H, Wurman J, Kosiba K (2017) Doppler radar-derived wind field of five tornado events with application to engineering simulations. Eng Str 148:509–521

    Article  Google Scholar 

  • Rosen KH (2000) Handbook of discrete and combinatorial mathematics. CRC Press, Boca Raton

    Google Scholar 

  • Savory E, Parke GAR, Zeinoddini M, Toy N, Disney P (2001) Modelling of tornado and microburst-induced wind loading and failure of a lattice transmission tower. Eng Str 23(4):365–375

    Article  Google Scholar 

  • Shehata AY, El Damatty AA, Savory E (2005) Finite element modeling of transmission line under downburst wind loading. Finite Elem Anal Des 42:71–89

    Article  Google Scholar 

  • Strawn RC, Kenwright DN, Ahmad J (1999) Computer visualization of vortex wake systems. AIAA J 37(4):511–512

    Article  Google Scholar 

  • Sujudi D, Haimes R (1995) Identification of swirling flow in 3-D vector fields. In: Proceedings of the 12th computational fluid dynamics conference, fluid dynamics and co-located conferences, San Diego, USA A95–36567

  • Tang Z, Feng C, Wu L, Zuo D, James DL (2018) Characteristics of tornado-like vortices simulated in a large-scale ward-type simulator. Boundary-Layer Meteorol 166(2):327–350

    Article  Google Scholar 

  • Ward NB (1972) The exploration of certain features of tornado dynamics using a laboratory model. J Atmos Sci 29(6):1194–1204

    Article  Google Scholar 

  • Wan CA, Chang CC (1972) Measurement of the velocity field in a simulated tornado-like vortex using a three-dimensional velocity probe. J Atmos Sci 29:116–127

    Article  Google Scholar 

  • Wang J, Cao S, Pang W, Cao J (2017) Experimental study on effects of ground roughness on flow characteristics of tornado-like vortices. Boundary-Layer Meteorol 162(2):319–339

    Article  Google Scholar 

  • Wen Y-K (1975) Dynamic tornadic wind loads on tall buildings. J Str Div 101:169–185

    Article  Google Scholar 

  • Wen Y-K, Ang AH-S (1975) Tornado risk and wind effects on structures. In: Eaton KJ (eds) Proceedings of the 4th international conference on wind effects on buildings and structures, London, UK, pp 63–74

  • Wong KY, Yip CL (2009) Identifying centers of circulating and spiraling vector field patterns and its applications. Ptrn Recognit 42(7):1371–1387

    Article  Google Scholar 

  • Wood VT, Brown RA (1992) Effects of radar proximity on single-Doppler velocity signatures of axisymmetric rotation and divergence. Mon Weather Rev 120(12):2798–2807

    Article  Google Scholar 

Download references

Acknowledgements

The authors should like to thank the Natural Sciences and Engineering Research Council (NSERC) of Canada and the University of Western Ontario for their financial support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Niall Bannigan.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1: Summary of Processing Algorithm

The centre of a vortex may be tracked with the proposed method by performing the following series of calculations on a two-dimensional, horizontal slice of data from a tornado dataset. Here, each layer is analyzed without consideration of the layers above or below (Fig. 6a). If not already available in the dataset, the vorticity, \({\varvec{\upomega}}\), is computed for the two-dimensional plane:

$${\varvec{\upomega}}=\nabla \times \mathbf{u}=\left(\frac{\partial {u}_{z}}{\partial y}-\frac{\partial {u}_{y}}{\partial z}\right)\widehat{\mathbf{i}}+\left(\frac{\partial {u}_{x}}{\partial z}-\frac{\partial {u}_{z}}{\partial x}\right)\widehat{\mathbf{j}}+\left(\frac{\partial {u}_{y}}{\partial x}-\frac{\partial {u}_{x}}{\partial y}\right)\widehat{\mathbf{k}}\boldsymbol{ },$$

where \(\mathbf{u}={u}_{x}\widehat{\mathbf{i}}+{u}_{y}\widehat{\mathbf{j}}+{u}_{z}\widehat{\mathbf{k}}\) is the wind-field velocity vector where \({u}_{x}\), \({u}_{y}\), and \({u}_{z}\) are the wind speeds in the x-, y-, and z-directions, respectively, but only the \(\widehat{\mathbf{k}}\) remains. Then, to ensure that the analysis does not proceed with too many datapoints for each step, a scale is determined based on the resolution and dimensions of the dataset using

$$ n = \frac{{\min \lfloor {n_{x} ,n_{y} } \rceil }}{c}, $$

where \(n\) is the scaling-factor (rounded to the nearest integer) based on the size of the dataset, \({n}_{x}\) is the number of points in x, \({n}_{y}\) is the number of points in y, and \(c\) is the scaling parameter (see Appendix 2).

The first \(n\) largest curl points (Fig. 6b), [\({x}_{\mathrm{c}},{y}_{\mathrm{c}}\)], are spatially averaged using

$$\langle {x}_{\mathrm{c}}\rangle =\frac{{\sum }_{i=1}^{n}{x}_{\mathrm{c},i}}{n} ,$$

and

$$\langle {y}_{\mathrm{c}}\rangle =\frac{{\sum }_{i=1}^{n}{y}_{\mathrm{c},i}}{n} .$$

The standard deviation, \(\sigma \), of these locations, in \(x\) and \(y\), respectively, is calculated using

$${\sigma }_{x}=\lceil\sqrt{\frac{\sum_{i=1}^{n}{\left({x}_{\mathrm{c}, i}-\langle {x}_{\mathrm{c}}\rangle \right)}^{2}}{n-1}} \rceil ,$$

and

$${\sigma }_{y}=\lceil\sqrt{\frac{\sum_{i=1}^{n}{\left({y}_{\mathrm{c},i}-\langle {y}_{\mathrm{c}}\rangle \right)}^{2}}{n-1}} \rceil .$$

The values are rounded up to accommodate that the data locations are not continuous and also to take a more conservative selection approach.

Then, a desired number of standard deviations, \(s\), from the averaged position is selected such that all of the maximum curl locations outside of the bounds of the ellipse, drawn by the following curve, are excluded:

$$\frac{{\left({x}_{\mathrm{c},i}-\langle {x}_{\mathrm{c}}\rangle \right)}^{2}}{{\left({\sigma }_{x}\right)}^{2}}+\frac{{\left({y}_{\mathrm{c},i}-\langle {y}_{\mathrm{c}}\rangle \right)}^{2}}{{\left({\sigma }_{y}\right)}^{2}}\le s .$$

The remaining number of positions, \(N\), will be used as the starting point of lines drawn perpendicular to the vector direction of the datapoint at each location (Fig. 6c). The location of each intersection, [\({x}_{\mathrm{t}},{y}_{\mathrm{t}}\)], between each line is computed using

$${x}_{\mathrm{t}}=\frac{{b}_{j}-{b}_{i}}{{m}_{i}-{m}_{j}} ,$$

and

$${y}_{\mathrm{t}}=\frac{{m}_{j}{b}_{i}-{m}_{i}{b}_{j}}{{m}_{j}-{m}_{i}} ,$$

where \(b\) is the \(y\)-intercept of the lines with the domain ordinate, \(m\) is the slope of the lines, and i and j are referring to the indexing of each line being compared such that i varies from 1 to \(N-1\) and j from \(i+1\) to N (Fig. 6d). The definition of \(N\) is modified to encompass only the intersections that are located within the domain range. The location of [\(\langle {x}_{\mathrm{t}}\rangle ,\langle {y}_{\mathrm{t}}\rangle \)] is then computed using

$$\langle {x}_{\mathrm{t}}\rangle =\frac{{\sum }_{i=1}^{N}{x}_{\mathrm{t},i}}{N} ,$$

and

$$\langle {y}_{\mathrm{t}}\rangle =\frac{{\sum }_{i=1}^{N}{y}_{\mathrm{t},i}}{N} .$$

The standard deviation, \(\sigma \), of these locations, in \(x\) and \(y\), respectively, is calculated using

$${\sigma }_{x \mathrm{t}}=\lceil\sqrt{\frac{\sum_{i=1}^{N}{\left({x}_{\mathrm{t}, i}-\langle {x}_{\mathrm{t}}\rangle \right)}^{2}}{N-1}} \rceil ,$$

and

$${\sigma }_{y \mathrm{t}}=\lceil\sqrt{\frac{\sum_{i=1}^{N}{\left({y}_{\mathrm{t},i}-\langle {y}_{\mathrm{t}}\rangle \right)}^{2}}{N-1}} \rceil .$$

Then, a desired number of standard deviations, \(s\) (its value may differ from that previously used), from the averaged position is selected such that all of the intersections outside of the bounds of the ellipse, drawn by the following curve, are excluded:

$$\frac{{\left({x}_{\mathrm{t},i}-\langle {x}_{\mathrm{c}}\rangle \right)}^{2}}{{\left({\sigma }_{x \mathrm{t}}\right)}^{2}}+\frac{{\left({y}_{\mathrm{t},i}-\langle {y}_{\mathrm{t}}\rangle \right)}^{2}}{{\left({\sigma }_{y \mathrm{t}}\right)}^{2}}\le s .$$

From the remaining number of intersections, \(I\), an average location is computed using

$$\langle {x}_{\mathrm{I}}\rangle =\frac{{\sum }_{i=1}^{I}{x}_{\mathrm{t},i}}{I} ,$$

and

$$\langle {y}_{\mathrm{I}}\rangle =\frac{{\sum }_{i=1}^{I}{y}_{\mathrm{t},i}}{I} .$$

This final location (Fig. 6e) should be a valid position and also sufficiently far from the domain edge such that the subsequent centre search does not attempt to collect data from outside of the domain. In the event that this does occur, the maximum \(z\)-vorticity or minimum pressure location can be substituted in place of [\(\langle {x}_{\mathrm{I}}\rangle ,\langle {y}_{\mathrm{I}}\rangle \)].

The nearest indexed position to [\(\langle {x}_{\mathrm{I}}\rangle ,\langle {y}_{\mathrm{I}}\rangle \)] is used as the centre of a grid of size \(c\)-by-\(c\) that will be used to identify the tornado centre, defined where there is zero horizontal, storm-relative velocity. The location of minimum velocity of these points is identified and, unless it is actually the tornado centre, is then used as the centre of a new smaller grid of size three-by-three. Each possible combination of loop pattern is identified so that there can be four quadrants made up of four positions in the new grid, four quadrants of six positions in the new grid, and a large loop around the outside of the new grid (see Fig. 7). The values of the signs of the vectors, \(S\), in each loop are summated so that only the loops where this sum is equal to zero are noted as being the possible location of the tornado centre:

$$\sum_{i=1}^{n}{S}_{i}=0 .$$

The sum of the magnitude of the tangential velocity vector, \({u}_{t}\), around each of these identified loops is calculated so that the one with the minimum \(\left|{u}_{t}\right|\) sum is taken to be the one with the tornado centre (Fig. 6f):

$$\mathrm{min}\sum_{i=1}^{n}{\left|{u}_{t}\right|}_{i} .$$

If the smaller grid is still too large, the above steps are applied iteratively until the grid is sufficiently small. If there is no path successfully identified, then the above steps are repeated, but the previously identified minimum velocity vector is skipped. The number of repetitions should be limited to a total of \(c\). Should this still fail to yield a path, the point used for the final averaged intersection location may be substituted as the tornado centre.

The four corners of the loop found above are bilinearly interpolated for the position where \({u}_{x}=0\) and \({u}_{y}=0\) (see Fig. 10 a), unless no path is found in which case this interpolation step is omitted. These corner points are arranged such that \(x\left(1\right)=x\left(3\right)\), \(x\left(2\right)=x\left(4\right)\), \(y\left(1\right)=y\left(2\right)\), and \(y\left(3\right)=y\left(4\right)\) and such that \(x\left(1\right)<x(2)\) and \(y\left(1\right)>y(3)\) (Fig. 10b). The process of the bilinear interpolation is given for the first step in finding the interpolation of the x-position between points 1 and 2 by

$${x}_{12}=x\left(1\right)+\left[\frac{0-{u}_{y,1}}{{u}_{y,2}-{u}_{y,1}}\right]\left(x\left(2\right)-x\left(1\right)\right) ,$$

and 3 and 4 by

$${x}_{34}=x\left(3\right)+\left[\frac{0-{u}_{y,3}}{{u}_{y,4}-{u}_{y,3}}\right]\left(x\left(4\right)-x\left(3\right)\right) ,$$

for zero velocity in the y-direction, and then the interpolation of the y-position between points 1 and 3 by

$${y}_{13}=y\left(1\right)+\left[\frac{0-{u}_{x,1}}{{u}_{x,3}-{u}_{x,1}}\right]\left(y\left(3\right)-y\left(1\right)\right) ,$$

and 2 and 4 by

$${y}_{24}=y\left(2\right)+\left[\frac{0-{u}_{x,2}}{{u}_{x,4}-{u}_{x,2}}\right]\left(y\left(4\right)-y\left(2\right)\right) ,$$

for zero velocity in the x-direction (Fig. 10c). Next, the velocities of these interpolated points are calculated for use in the final interpolation of the position of zero tangential velocity that identifies the centre of the tornado. The velocity in the x-direction between points 1 and 2 is interpolated using

$${u}_{x,12}={u}_{x,1}+\left[\frac{{x}_{12}-x\left(1\right)}{x(2)-x(1)}\right]\left({u}_{x,2}-{u}_{x,1}\right) ,$$

then in the x-direction between points 3 and 4 using

$${u}_{x,34}={u}_{x,3}+\left[\frac{{x}_{34}-x\left(3\right)}{x(4)-x(3)}\right]\left({u}_{x,4}-{u}_{x,3}\right) ,$$

in the y-direction between points 1 and 3 using

$${u}_{y,13}={u}_{y,1}+\left[\frac{{y}_{13}-y\left(1\right)}{y(2)-y(1)}\right]\left({u}_{y,3}-{u}_{y,1}\right) ,$$

and finally, in the y-direction between points 2 and 4 using (Fig. 10d)

$${u}_{y,24}={u}_{y,2}+\left[\frac{{y}_{24}-y\left(2\right)}{y(4)-y(2)}\right]\left({u}_{y,4}-{u}_{y,2}\right) .$$
Fig. 10
figure 10

Finding the tornado centre using artificial data for illustrative purposes as an example. a Schematic example of tornado dataset with vectors (shown as black arrows) on an isotropic grid (shown as dashed lines), b vector components (\(x\) in green and \(y\) in blue), c location of zero velocity (between \(x\) components in green and \(y\) components in blue), d interpolated vector components, e location of zero horizontal velocity (shown as red circle), and f tornado centre location

Hence, the x and y positions of zero tangential velocity between these calculated velocities can be interpolated to identify the centre of the tornado (Fig. 10e)

$$x={x}_{12}+\left[\frac{0-{u}_{x,12}}{{u}_{x,34}-{u}_{x,12}}\right]\left({x}_{34}-{x}_{12}\right) ,$$

and

$$y={y}_{13}+\left[\frac{0-{u}_{y,13}}{{u}_{y,24}-{u}_{y,13}}\right]\left({y}_{24}-{y}_{13}\right) ,$$

which concludes the searching portion of the algorithm (Fig. 10f).

Appendix 2: Comparison of Scale Factor, c

To test the effects of changing the number of points selected in the maximum curl identification process, c was varied from 1 to 10, as an integer only for simplicity, and, based on trial-and-error, was found to be an appropriate range. Although the overall effect of changing this value appears to be minimal, selecting an extreme value in this range will quickly deteriorate the accuracy of the centre detection capabilities of the proposed method and increase the chance of another smaller, less-defined vortex being included in the search (see Table

Table 3 Comparison of the number of times the programme failed to find a loop about a centre and/or interpolate for a centre using different c values

3).

An optimum value of c = 6 was determined for the current data based on the least number of errors and re-attempts, even though a non-integer value may be optimal. The paths determined for each different c-value can be seen in Fig. 

Fig. 11
figure 11

Comparing the mode of identified tornado centre points, found by varying c, with the failed results of the different c values used. The data shown are taken relative to the simulation domain for all timesteps at height z = 15 m a.g.l

11. Adjusting c appears to only affect the searching algorithm negatively at either end of the range selected. This is because it begins to either exclude so many points that almost none are left for the analysis or, at the other extreme, include so many points that the selection criterion of maximum curl becomes meaningless. This scaling factor acts as a tool to eliminate the many data points identified as having a high curl but that only represent partially formed or small vortices and would otherwise skew the centre locating of the program. It should only need to be adjusted once by the user for a given dataset before proceeding with further analyses.

The searches performed neglected the results of the search from each preceding timestep, resulting in errors introduced by sweeping through the entire domain unnecessarily. Given the high spatial and temporal resolution of the dataset used in this paper, the search domain after the centre is successfully located the first time may be reduced to a very small region and carried through to each subsequent timestep processed.

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bannigan, N., Orf, L. & Savory, E. Tracking the Centre of Asymmetric Vortices Using Wind Velocity Vector Data Fields. Boundary-Layer Meteorol 186, 1–26 (2023). https://doi.org/10.1007/s10546-022-00739-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10546-022-00739-0

Keywords

Navigation