Paper The following article is Open access

Modeling cloud-to-ground lightning probability in Alaskan tundra through the integration of Weather Research and Forecast (WRF) model and machine learning method

and

Published 2 November 2020 © 2020 The Author(s). Published by IOP Publishing Ltd
, , Resiliency and Vulnerability of Arctic and Boreal Ecosystems to Environmental Change: Advances and Outcomes of ABoVE (the Arctic Boreal Vulnerability Experiment) Citation Jiaying He and Tatiana V Loboda 2020 Environ. Res. Lett. 15 115009 DOI 10.1088/1748-9326/abbc3b

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1748-9326/15/11/115009

Abstract

Wildland fires exert substantial impacts on tundra ecosystems of the high northern latitudes (HNL), ranging from biogeochemical impact on climate system to habitat suitability for various species. Cloud-to-ground (CG) lightning is the primary ignition source of wildfires. It is critical to understand mechanisms and factors driving lightning strikes in this cold, treeless environment to support operational modeling and forecasting of fire activity. Existing studies on lightning strikes primarily focus on Alaskan and Canadian boreal forests where land-atmospheric interactions are different and, thus, not likely to represent tundra conditions. In this study, we designed an empirical-dynamical method integrating Weather Research and Forecast (WRF) simulation and machine learning algorithm to model the probability of lightning strikes across Alaskan tundra between 2001 and 2017. We recommended using Thompson 2-moment and Mellor–Yamada–Janjic schemes as microphysics and planetary boundary layer parameterizations for WRF simulations in the tundra. Our modeling and forecasting test results have shown a strong capability to predict CG lightning probability in Alaskan tundra, with the values of area under the receiver operator characteristics curves above 0.9. We found that parcel lifted index and vertical profiles of atmospheric variables, including geopotential height, dew point temperature, relative humidity, and velocity speed, important in predicting lightning occurrence, suggesting the key role of convection in lightning formation in the tundra. Our method can be applied to data-scarce regions and support future studies of fire potential in the HNL.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 license. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Wildland fire is a dominant disturbance across boreal forest and tundra ecosystems in the high northern latitudes (HNL; Goetz et al 2005, Bond-Lamberty et al 2007, French et al 2015). Tundra fires can alter ecosystem functioning and drive environmental changes in carbon cycling and energy budget (Mack et al 2011, Bret-Harte et al 2013, Pearson et al 2013, Jones et al 2015, French et al 2016). Specifically, they can release large stores of carbon locked in organic soil and permafrost. For example, the 2007 Anaktuvuk River fire in Alaska burned 1039 km2 and released ∼2.1 Tg carbon into the atmosphere (Mack et al 2011). Although much rarer and generally less severe than boreal fires, they are common, particularly in Alaska. Between 2001 and 2015, Alaskan tundra has burnt about 54% of 10 260 km2 burned area in the tundra worldwide (He et al 2019). Additionally, post-fire recovery for critical components of the tundra ecosystem can last for several decades, which is comparable to that of the boreal forests (Racine et al 1987, Jandt et al 2008). Thus, tundra fires can impact long-term winter forage availability, particularly lichen, for caribou, which will subsequently influence the subsistence resources of local human communities (Gustine et al 2014, Bliss and Hu 2020).

Fires in the remote and generally inaccessible HNL are primarily ignited by cloud-to-ground (CG) lightning flashes (French et al 2015, Veraverbeke et al 2017), also known as lightning strikes. Future projections indicate potential increase of lightning under climate warming (Price and Rind 1994a, Romps et al 2014), which will subsequently lead to more fire occurrences and larger burned areas in the HNL (Wotton et al 2010, Krause et al 2014, French et al 2015, Veraverbeke et al 2017). Numerous studies have explored the driving factors of lightning strikes and developed predictive models (Burrows et al 2005, Blouin et al 2016) for Alaskan and Canadian boreal forests. However, considerably less is known about factors driving CG lightning for the treeless tundra. The substantial differences in surface conditions between tree-dominated and treeless landscapes (Dissing and Verbyla 2003, Beringer et al 2005, Van Heerwaarden and Teuling 2014, Rivas Soriano et al 2019) imply that understanding lightning processes in the boreal forests is not necessarily readily transferable to the tundra. Therefore, tundra-focused studies are critical for enhancing the modeling capability of lightning and fire potential, and assisting wildfire management efforts in the future.

Typically, lightning formation is associated with atmospheric convection in cumulonimbus clouds (Anderson 1992). Lightning flashes are generated through buildup, separation, and transfer procedures of electric charges between cloud particles (Saunders 2008, Yair 2008). The occurrence and intensity of lightning are generally related to factors such as convective cloud development, cloud structure, and hydrometeor attributes (Price and Rind 1994b, Baker et al 1999, Buiat et al 2017). However, explicit simulation and prediction of the electrification processes can be computationally expensive (Zepka et al 2014). Further efforts are also required to comprehensively understand the detailed microphysical procedures contributing to the charge accumulation (Rakov and Uman 2003, Saunders 2008). Efforts on modeling lightning activity thus primarily rely on developing relationships with observed or model-resolved parameters related to convective activities and cloud microphysical properties.

Classificatory schemes or regression methods developed with convective indices or weather conditions were among the earliest lightning modeling attempts (Anderson 1991, Sly 1965, Reap and Foster 1979, Fuquay 1980, Reap and Macgorman 1989). With the development of numerical weather prediction (NWP) models, lightning schemes based on microphysics principles have been parameterized within models from regional to global scales (Price and Rind 1994b, Barthe et al 2010, Yair et al 2010, Lynn et al 2012, Wong et al 2013). Simple strategies were also designed with Weather Research and Forecast (WRF) simulations (Zepka et al 2014, Giannaros et al 2015). In addition to physical parameterizations, empirical-based methods such as logistic regression (Shafer and Fuelberg 2006, Bates et al 2018) and random forest (RF; Blouin et al 2016, Schön et al 2018) were adopted to model lightning based on meteorological conditions and thunderstorm characteristics. Opportunities for integrating dynamic NWP and statistical models have been explored to improve the modeling capability of lightning potential. Burrows et al( 2005) trained tree-structured regression models to forecast lightning probability based on the Global Environmental Multiscale (GEM) model. Sousa et al( 2013) and Gijben et al( 2017) also combined WRF with logistic regression to develop a statistical-dynamical method for CG lightning prediction in different regions.

Due to the remoteness of tundra, meteorological observations are very sparsely distributed and thus unsuitable for describing the spatial variation of tundra conditions. Although reanalysis products ensure spatial-temporal consistency, their performances are limited by the coarse resolution, limited availability of observations, and uncertainty of diagnostic variables (Dee et al 2016). Therefore, purely empirical models trained with these data are inappropriate for lightning modeling in data-scarce regions like tundra. Although NWP has not been specifically adopted in tundra studies, existing research has demonstrated its suitability and effectiveness for modeling lightning potential (Reap 1991, Burrows et al 2005) and fire danger (Mölders 2010, Di Giuseppe et al 2016) in boreal forests. In this study, we aim to understand what atmospheric factors affect the CG lightning probability in Alaskan tundra from 2001 to 2017 and to model the probability by integrating WRF and RF.

2. Study area

We defined Alaskan tundra using the Circumpolar Arctic Vegetation Map (CAVM; figure 1(a); Walker et al 2005). Fire regimes vary by year and across different tundra regions (figure 2). In Alaska, more than 99% of the lightning strikes occur from May to August (Reap 1991, Mcguiney et al 2005). Lightning activity shows a typical diurnal pattern from noon until midnight with a peak between 4 pm and 8 pm. Elevation and forest cover can affect the spatial variation of lightning strikes in Alaska by altering the convective activity (Dissing and Verbyla 2003). In particular, large-scale atmospheric instability and local convergence are the primary contributors to the formation thunderstorm of lightning (Reap 1991).

Figure 1.

Figure 1. Study area: (a) Alaskan tundra defined by CAVM; (b) domain configuration for two-way nested WRF simulation.

Standard image High-resolution image
Figure 2.

Figure 2. Fire count in different regions of Alaskan tundra and total burned area for each year from 2001 to 2017 according to Alaska Large Fire Database (ALFD).

Standard image High-resolution image

3. Materials and methods

We designed an empirical-dynamical method using WRF-simulated atmospheric variables to model CG lightning probability and understand its driving factors in Alaskan tundra. We first conducted a sensitivity analysis to identify optimal WRF parameterizations that best describes tundra meteorological conditions. We then ran WRF and developed RF models for predicting CG lightning probability.

3.1. CG lighting observations

We obtained lightning data from the Alaska Lightning Detection Network (ALDN; Fronterhouse 2012), with a detection efficiency better than 5 km and positional accuracy higher than 70% (Reap 1991, Dissing and Verbyla 2003). The system has been updated to improve detection performance (Fronterhouse 2012). Devices employed before 2012, developed by Vaisala, Inc., recorded lightning flash with multiplicity (i.e., lightning strikes per flash). The new system, provided by TOA Systems, Inc., records lightning strikes instead (Fronterhouse 2012). To ensure data consistency between systems, we adopted lightning strikes rather than flashes in this study.

3.2. WRF setup and sensitivity analysis

We adopted the Advanced Research WRF version 4.0 (Skamarock et al 2019) for simulation. The National Centers for Environmental Prediction Final Operational Model Global Tropospheric Analysis data (NCEP FNL) at 1° × 1° resolution and 6-h interval were used for model initialization (NCEP 2000). We defined two domains with 25 km (Domain 1) and 5 km (Domain 2) grid spacings for two-way nested simulation (figure 1(b)). The vertical dimension was configured with 33 unevenly spaced full sigma levels with the model top defined at 50 hPa.

WRF provides multiple parameterization schemes for major physics components (table 1), which can affect simulation results for a specific region given various assumptions and mechanisms. Since existing WRF applications in the HNL primarily focus on the boreal forests or the pan-Arctic region, their schemes may not be suitable for the tundra. Therefore, we conducted a sensitivity analysis to determine WRF settings that most closely match tundra meteorological observations.

Table 1. WRF parameterization schemes used for the pan-Arctic region in existing literature.

ComponentsSchemesReferences
Microphysics Morrison 2-moment (Bieniek et al 2016)
Thompson 2-moment (Kim et al 2014; Mölders 2010, Mölders 2008)
WRF single-moment 6-class (Hines and Bromwich 2008)
Goddard Cumulus Ensemble Model (Cassano et al 2011; Glisan et al 2013)
Longwave radiation Rapid Radiative Transfer Model (RRTM) Longwave (Bieniek et al 2016; Cai et al 2018; Glisan et al 2013; Hines et al 2011; Kim et al 2014; Mölders 2010, Mölders 2008)
Community Atmospheric Model (CAM) Longwave (Cassano et al 2011; Glisan et al 2013)
Shortwave radiation RRTM Shortwave (Bieniek et al 2016; Glisan et al 2013; Kim et al 2014)
Dubhia Shortwave (Cai et al 2018; Mölders 2010, Mölders 2008)
Goddard Shortwave (Hines et al 2011)
CAM Shortwave (Cassano et al 2011; Glisan et al 2013)
Cumulus Grell-Devenyi ensemble (Bieniek et al 2016; Cassano et al 2011; Glisan et al 2013)
Kain-Fritsch (Cai et al 2018)
PBL/ Surface layer Mellor-Yamada-Janjic (MYJ)/Monin-Obukhov Eta (Eta) (Bieniek et al 2016; Cassano et al 2011; Glisan et al 2013; Hines et al 2011)
Yonsei University (YSU)/ Monin-Obukhov MM5 (MM5) (Cai et al 2018; Kim et al 2014; Mölders 2010, Mölders 2008)
Land surface model Noah (Bieniek et al 2016; Cai et al 2018; Cassano et al 2011; Glisan et al 2013; Hines et al 2011)
Rapid Update Cycle (Mölders 2010, Mölders 2008)

We first identified a list of candidate schemes through literature review, the majority of which utilized the Noah, Rapid Radiative Transfer Model (RRTM), and Grell–Devenyi schemes for land surface model, radiation, and cumulus components, respectively (table 1). We adopted their updated versions in WRF version 4.0. Mixed-phased microphysics schemes are typically recommended for simulating icing or convective conditions at a horizontal resolution finer than 10 km (Skamarock et al 2019). We thus selected 2-moment schemes of Morrison, Thompson, and WRF 6-class (WRF6) for microphysics, and considered both Mellor–Yamada–Janjic (MYJ) and Yonsei University (YSU) schemes for planetary boundary layer (PBL) representation in our sensitivity analysis (table 2).

Table 2. Six candidates of parameterization combinations for sensitivity analysis.

Combination notationMicrophysicsPBL + surface layer
Morrison_MYJ Morrison 2-moment MYJ + Eta
Thompson_MYJ Thompson 2-moment MYJ + Eta
WRF6_MYJ WRF6 MYJ + Eta
Morrison_YSU Morrison 2-moment YSU + MM5
Thompson_YSU Thompson 2-moment YSU + MM5
WRF6_YSU WRF6 YSU + MM5

Remote Automated Weather Stations (RAWS; https://raws.nifc.gov) provide the densest network of near-surface weather observations across Alaska to date. Four variables, including air temperature (T), dew point temperature (Td), relative humidity (RH), and solar radiation, were utilized as 'ground truth' for sensitivity analysis. To consider different fire regimes (figure 2), we selected four cases for 24 h simulations in years of varying fire season intensity (2006—low, 2007—moderate, 2010 and 2015—intense, figure 2; table 3). We calculated three statistical metrics for each variable, including root-mean-square error (RMSE), mean absolute error (MAE) and Pearson's r correlation, and ranked them of all candidates for each year from 1 (lowest) to 6 (highest). For each metric, the yearly rankings were summed up to create a single rank sum (ranging from 4 to 24), with the larger value representing better overall performance.

Table 3. Summary of case studies for sensitivity analysis.

YearNumber of RAWS stationsSimulation period (UTC time)
2006 8 08/16 00:00–08/17 00:00
2007 9 05/01 00:00–05/02 00:00
2010 12 07/01 00:00–07/02 00:00
2015 29 06/18 00:00–06/19 00:00

3.3. Random forest modeling of CG lightning probability

We used the RF classification algorithm (Breiman 2001) to model CG lightning probability with WRF-simulated atmospheric variables as predictors. As an ensemble method, RF generates a large number of decision trees through permutation and integrates their results for a more stable modeling performance. Two separate modeling experiments referred to as '24 hr model' and '48 hr model' were designed to compare the consistency of modeling performance and variable importance with different simulation periods (0–24 h and 0–48 h, respectively; figure 3).

Figure 3.

Figure 3. Illustration of the '24 h model' and '48 h model' components.

Standard image High-resolution image

For each experiment, atmospheric variables were extracted from WRF outputs (Domain 2) simulated at 24 h or 48 h after initialization, respectively. Four groups of variables were summarized from literature (Reap 1991, Burrows et al 2005, Sousa et al 2013, Blouin et al 2016), including stability indices, cloud properties, weather conditions at multiple pressure levels (500 hPa, 700 hPa, 850 hPa, and 1000 hPa), and two lightning parameterizations from WRF (table 4). Then we used these predictors to model the presence and absence of lightning strikes during the following 24 h after the timing of variable extraction (figure 3). Lightning points from the ALDN dataset were labeled as presence, while random sample points within tundra with no lightning occurred were labeled as absence.

Table 4. List of independent variables retrieved from WRF simulation output for RF modeling.

CategoriesAbbreviationDescription
Atmospheric stability indices CAPE Convective Available Potential Energy
CIN Convective Inhibition
LCL Lifted Condensation Level
LFC Level of Free Convection
TT Total Totals
KI K Index
PLI Parcel Lifted Index (to 500 hPa)
BI Boyden Index
SHOW Showalter Index
Cloud properties CFtotal Total cloud cover fraction
CFhigh High-level cloud cover fraction
CFmid Mid-level cloud cover fraction
CF low Low-level cloud cover fraction
CTT Cloud top temperature
CTH Cloud top height
CTP Cloud top pressure
IWP Ice water path
LWP Liquid water path
ERIce Effective radius of cloud ice
ERWater Effective radius of cloud water
QCloud Cloud water mixing ratio
QIce Ice mixing ratio
QRain Rain mixing ratio
BT Brightness temperature
Weather variables T Air temperature at surface and multiple pressure levels
Td Dewpoint temperature at surface and multiple pressure levels
T.Td Temperature-dewpoint spread at multiple pressure levels
RH Relative humidity at surface and multiple pressure levels
GPZ Geopotential height at multiple pressure levels
DZ Thickness between any two pressure layers
W Vertical velocity at multiple pressure levels
Helicity Helicity
UH Updraft helicity
Rain Total precipitation
PW Precipitable water
SLP Sea level pressure
Lightning parameterizations PR92 Flash distribution of CG lightning with PR92 scheme
LPI Lightning probability index

To ensure the representativeness of our models in describing tundra lightning, three lightning severity levels were identified based on the total number of daily lightning strikes for model training, following a similar method by Farukh et al( 2011): >2000 strikes per day as severe, 500–2000 strikes per day as moderate, and 0–500 strikes per day as low. For each level, five cases were selected for WRF simulation using a stratified random sampling strategy (table 5). We then randomly selected 70% of presence and absence points for model training and reserved the rest 30% for validation. Training and testing points from all levels were combined for model development and accuracy assessment, respectively. We set the number of trees to 500 and the number of variables at each split as 8 in the RF algorithm.

Table 5. Cases studies of WRF simulations for CG lightning modeling.

Initialization ofTotal CG lightning strokesTotal CG lightning strokes
simulation (UTC)during 0 ∼ 24 hduring 24 ∼ 48 h
2003/06/23 00:00 1305 1560
2005/06/11 00:00 732 1027
2005/06/29 00:00 3449 2151
2005/08/16 00:00 5 15
2007/07/04 00:00 4280 479
2007/07/11 00:00 2321 2247
2008/06/25 00:00 2030 295
2009/06/09 00:00 335 64
2010/07/01 00:00 1015 1991
2013/06/20 00:00 3073 3641
2013/08/16 00:00 60 10
2015/07/14 00:00 847 4320
2015/06/21 00:00 974 905
2015/07/23 00:00 175 80
2016/07/11 00:00 93 109

We reported the out-of-bag (OOB) error rate to evaluate the overall accuracy of model training. For validation, we calculated statistical criteria based on the reserved dataset (tables 6 and 7), and generated receiver operating characteristics (ROC) curves and area under the curve (AUC). Moreover, we examined the contribution of predictors in modeling tundra lightning potential with variable importance quantified by mean decrease in accuracy (MDA).

Table 6. Contingency matrix of variables used to calculate statistical scores.

  Observed
CG lightning eventPresenceAbsence
Predicted Presence $a$ (hit) $b$ (false alarm)
Absence $c$ (miss) $d$ (correct non-event)

Table 7. Statistical criteria used for assessing modeling performance.

Statistical scoresAbbreviationFormula
Probability of Detection POD $a/\left( {a + c} \right)$
Critical Success Index CSI $a/\left( {a + b + c} \right)$
False Alarm Ratio FAR $b/\left( {a + c} \right)$
False Alarm Rate F $b/\left( {b + d} \right)$

3.4. Forecasting capability assessment

In addition to empirical modeling, we tested the capability of our empirical-dynamical method in forward forecasting of CG lightning probability at a future timing by applying the RF model developed for a previous period. Here we utilized the '24 hr model' to forecast the CG lightning probability with atmospheric conditions simulated 48 h after initialization. Statistical criteria listed in table 7 and ROC curves were also generated to quantify the forecasting capability.

4. Results

4.1. Sensitivity analysis of WRF simulation

The results of sensitivity analysis indicate variability in the performance of different WRF schemes by meteorological variables and across cases (figure 4; Tables S1–S4 (available online at https://stacks.iop.org/ERL/15/115009/mmedia)). The combined statistical ranking for each variable can range between 4 (the lowest rank across all four sensitivity cases, see table 3) and 24 (the highest rank). Based on the results, Thompson_MYJ and Morrison_MYJ emerged as the strongest performing settings for tundra meteorology simulations (figure 4). Although the majority of the correlation values for all metrological variables are around 0.7 ∼ 0.8, the overall simulation results with all six candidates for T and Td outperform those for RH and solar radiation according to RMSE and MAE (tables S1–S4). Specifically, Thompson_MYJ shows the best results compared to other schemes for T and Td( figure 4(a)–(b); Tables S1–S2). For RH, MYJ for the PBL scheme delivers superior results compared to YSU, particularly when combined with the Morrison and Thompson schemes (figure 3(c)). While for solar radiation, the Morrison scheme outperforms other microphysics schemes, followed by Thompson.

Figure 4.

Figure 4. Bar plots summarizing the ranks of six candidate schemes based on Pearson's r correlation, MAE and RMSE for: (a) T, (b) Td, (c) RH, and (d) solar radiation. The largest rank sum value represents the best overall performance.

Standard image High-resolution image

Since lightning activity is largely related to convection, we further compared the Thompson_MYJ and Morrison_MYJ schemes by examining the spatial distribution of convective available potential energy (CAPE) as a representative. Considering that the atmospheric soundings were very sparse in the tundra, and the spatial distribution of CAPE from the two schemes was fairly close, we chose the Thompson_MYJ scheme for further simulation, since it shows a more detailed distribution of CAPE values for describing convective activities (figure 5). The optimized WRF schemes were summarized in table 8.

Figure 5.

Figure 5. Comparison of CAPE simulation in Domain 2 between (a) Morrison_MYJ and (b) Thompson_MYJ for the 2010 case as an example.

Standard image High-resolution image

Table 8. Optimized combination of physical parameterizations for WRF.

Physical componentParameterization schemeSetting option in WRF
Microphysics Thompson 2-moment mp_physics = 8
Cumulus Grell-Freitas ensemble cu_physics = 3
PBL MYJ bl_pbl_physics = 2
Surface layer Eta sf_sfclay_physics = 2
Land surface model Noah sf_surface_physics = 2
Longwave radiation RRTMG ra_lw_physics = 4
Shortwave radiation RRTMG ra_sw_physics = 4

4.2. Accuracy assessment of RF models

The overall OOB error rate and the class errors suggest high accuracy in predicting CG lightning strikes in Alaskan tundra. The '24 hr model' has an overall error rate of 4.64% and an accuracy of 95.36%. Specifically, the absence of CG lightning has a class error of 7.26%, while that of the presence is only 2.54% (table 9(a)). The overall error rate of the '48-hr model' is 6.81% and the accuracy reaches 93.19%. The class errors of the absence and presence of lightning events are 11.15% and 3.43%, respectively (table 9(b)).

Table 9. Confusion matrices for (a) '24 hr model' and (b) '48 hr model'.

(a) 24 h model
    CG lightning predictions  
Confusion matrix Absence Presence Class error
CG lightning observations Absence 7013 549 0.0726
Presence 241 9239 0.0254
(b) 48 h model
    CG lightning predictions  
Confusion matrix Absence Presence Class error
CG lightning observations Absence 4872 612 0.1115
Presence 243 6837 0.0343

Our validation results against the reserved dataset show high overall accuracy in predicting CG lightning strikes in both models by assessing the statistical metrics (table 10) and the ROC curves (figure 6). We also found a notably stronger performance during severe and moderate lightning days than the low severity level in both models. The '24 hr model' shows an overall probability of detection (POD) of 0.96 and a critical success index (CSI) of 0.91, while the false alarm ratio (FAR) and false alarm rate (F) are below 0.1. The AUC values are above 0.95 across lightning days of all severity levels. The '48 hr model' shows an overall POD value of 0.97 and a CSI of 0.88, while the FAR and F metrics are around 0.1. However, the POD and CSI values are below 0.5 for the low severity cases, much lower than those from the '24 hr model'. The AUC values reported from validation data are higher than 0.95 during severe and moderate lightning days while dropping to 0.87 for the low severity level days (table 10).

Figure 6.

Figure 6. ROC curves of validation results on lightning days with different severity levels: (a) 24 h model, (b) 48 h model.

Standard image High-resolution image

Table 10. Statistical criteria calculated using the validation data for: (a) 24 h model; (b) 48 h model.

ModelsMetricsAllSevere levelModerate levelLow level
24 h model POD 0.9628 0.9821 0.9456 0.7885
CSI 0.9065 0.9365 0.8569 0.7454
FAR 0.0606 0.0472 0.0987 0.0682
F 0.0858 0.1167 0.0794 0.0277
AUC 0.9869 0.9898 0.9818 0.9743
48 h model POD 0.9668 0.9701 0.9643 0.2273
CSI 0.8815 0.8808 0.8833 0.1829
FAR 0.0909 0.0946 0.0868 0.5161
F 0.1247 0.1291 0.1196 0.0366
AUC 0.9810 0.9810 0.9809 0.8751

In addition to the purely statistical accuracy assessment, we generated CG lightning probability maps across the entire Alaskan tundra to visually compare the predicted spatial patterns to observed lightning strike distribution (see an example for a severe lightning day in figure 7).

Figure 7.

Figure 7. (a) Observed CG lightning strikes during 2010/07/02 00:00 and 2010/07/03 00:00 UTC and (b) modeled CG lightning probability during the same period with the '24 hr model' in Alaskan tundra.

Standard image High-resolution image

4.3. Forecasting performance

The forecasting test using the '24 hr model' for the 48-h simulation demonstrates promising forecasting results with our method, with POD and CSI around 0.7 and two false ratios FAR and F below 0.08 (table 11). When separated by different severity levels of lightning days, the forecasting performance is consistent with previously reported results: the accuracy appears to be substantially higher for severe and moderate lightning conditions than low severity days. Similar patterns can be found from the ROC curves (figure 8), with the AUC values around 0.9 for the entire data and those from the moderate level cases (table 11).

Figure 8.

Figure 8. ROC curves of forecasting test on lightning days with different severity levels.

Standard image High-resolution image

Table 11. Statistical criteria calculated for the forecasting test.

MetricsAllSevere levelModerate levelLow level
POD 0.7063 0.6372 0.7707 0.1515
CSI 0.6660 0.5898 0.7397 0.1397
FAR 0.0788 0.1119 0.0516 0.2
F 0.0780 0.1024 0.0547 0.0132
AUC 0.9203 0.8766 0.9519 0.769

4.4. Evaluation of variable importance

To understand the predictors' roles in determining the CG lightning potential in the tundra, we examined the top 20 important variables ranked according to MDA from both the '24 hr model' and the '48 hr model' for comparison (figure 9). According to both models, the parcel lifted index (PLI) is found to be the most important variable in determining the accumulated lightning strikes among all the predictors. For the '24 hr model', weather variables at multiple pressure levels, including geopotential height (GPZ), Td, RH, and velocity speed, are also shown as highly important in determining CG lightning occurrence in the tundra (figure 9(a)). Also, cloud fraction, sea level pressure, helicity, lifted condensation level, and atmospheric stability indices, such as the Total Totals and the Showalter Index (SHOW), are among the top 20 important predictors. Although the variable ranking order of the '48 hr model' differs from that of the '24 hr model', the majority of the top 20 important variables remain the same. Only layer thickness between 700 hPa and 850 hPa levels (DZ700–850) and brightness temperature appear to play an important role in determining the lightning potential for the '48 hr model' but not in the '24 hr model'.

Figure 9.

Figure 9. Top 20 important variables ranked by MDA in (a) '24 hr model' and (b) '48 hr model'.

Standard image High-resolution image

5. Discussions

In general, this study successfully demonstrates the strong capability of the empirical-dynamicalframework in representing meteorological conditions that support CG lightning well in Alaskan tundra. By integrating WRF and RF algorithm, our method shows excellent performance in both modeling CG lightning strikes with 24 h and 48 h WRF simulations, and forward forecasting of lightning probability. This supports the effectiveness of the framework in accurate prediction of future CG lightning potential in data-scarce regions like the HNL.

Our results identify the PLI to 500 hPa as the most important factor in modeling lightning in Alaskan tundra. Similarly, Farukh et al( 2011) endorsed the lifted index for characterizing lightning in Alaska because of its sensitivity to evaluate upper air instability. This highlights the key role of the lift potential in providing sufficient convection to support lightning formation in the tundra, consistent with the results of Burrows et al( 2005) in the far west and north region of North America dominated by tundra. While for boreal forests, SHOW and CAPE were typically recognized as the top-ranking indicators of lightning potential (Burrows et al 2005, Blouin et al 2016). Compared to static indices like SHOW, PLI can improve forecasting capability, especially during the afternoon (Galway 1956). Though commonly considered a superior measure of instability, CAPE was not found to be very important here. Unlike CAPE that represents integrated parcel buoyancy of the troposphere, PLI is more of a measure of actual 'instability' since it represents the potential buoyancy of a parcel at a certain level (Blanchard 1998). The various index performance in these studies may reflect different land-atmosphere energy exchange and convection between tundra and boreal forests, which also affect the boundary conditions in these ecosystems (Thompson et al 2004, Beringer et al 2005).

Additionally, GPZ at 500 hPa ranks top in determining CG lightning distribution, indicating potential connections between this pressure level and lightning activity in Alaskan tundra. While Burrows et al( 2005) suggested that the occurrence of lightning was influenced by the interaction between strong convection and precipitable water (PW) in the cloud, here we found much lower rankings of PW than instability indices like PLI and SHOW in both models. This suggests that convection plays a more critical role than PW in determining the lightning potential in the tundra, which is consistent with the findings of Reap (1991). We also found that the vertical profiles of variables like GPZ, RH, Td, and layer thickness are highly important in modeling lightning activity in the tundra (figure 9). However, neither of the two internal WRF lighting parameterizations (PR92 and LPI) appears to represent lightning potential in the tundra at the regional scale well. This is not surprising as these parameterizations were not specifically developed for the HNL.

Our selection of Thompson 2-moment and MYJ schemes for WRF simulation in the tundra is consistent with the existing findings of WRF applications in lightning modeling. With a more detailed distribution of CAPE values for describing convective activities (figure 4), the Thompson scheme is also recommended by other studies given its detailed representation of ice-phase processes and improved simulation performance for convection related events like precipitation (Zepka et al 2014, Giannaros et al 2015). For lightning modeling purposes, the MYJ scheme is suitable for describing PBL conditions considering its optimal description of atmospheric conditions for triggering convection activities (Sousa et al 2013, Giannaros et al 2015).

Despite WRF's strong capabilities in regional modeling and dynamic downscaling of atmospheric conditions, its performance is affected by the parameterization schemes' assumptions and mechanisms for describing physical processes. Polar WRF is now under development to improve the representation of near-surface and atmospheric conditions for the Arctic regions, which can support lightning modeling for Alaska in the future (Hines et al 2011, Wilson et al 2011, Cai et al 2018).

Although our modeling results show very impressive prediction capabilities, they are subject to uncertainties inherent in CG lightning observations detected from either the ground-based networks. Although satellite-based sensors such as Lightning Imaging Sensor or Geostationary Lightning Mapper monitor lightning at a large spatial scale, their datasets are not available for the HNL due to the limited spatial coverage and resolution (Nag et al 2014, Matsangouras et al 2016). Separating the CG and cloud-to-cloud lightning flashes from satellite observations can introduce errors as well (Nag et al 2014). Although ground-based systems are constrained by position accuracy and detection efficiency, we found reporting probability a more effective way to describe CG lightning activity.

In addition to atmospheric factors that function as the key predictors for lightning, synoptic-scale dynamic forcings control meteorological mechanisms driving lightning and fire weather (Reap 1994, Flannigan and Wotton 2001, Santos et al 2013). For example, Kochtubajda et al( 2019) found more frequent ridging and ridge displacements during the 2014 wildfire season in the Northwest Territories of Canada. Synoptic weather conditions should be explored and incorporated in future modeling efforts to improve our understanding of lightning and fire regimes in the tundra.

Since CG lightning is the primary ignition source of wildfire in boreal forests and tundra (French et al 2015, Veraverbeke et al 2017), our integration of NWP and machine learning is easily transferable to monitor and predict fire potential in other data-scarce regions. This provides opportunities for fire management efforts by enhancing our capability to assess future fire impacts on ecosystems and climate and improving fire prevention and suppression practices. Also, fire prediction can ultimately improve the safety and health of the human communities as an important tool in public health applications. Wildfires in the HNL frequently occur in remote and sparsely populated areas and thus pose little direct threat to life and property. However, emissions can be transported hundreds to thousands of kilometers away, exerting considerable impacts on local and regional air quality and leading to various health outcomes (Morris et al 2006, Reisen et al 2015).

6. Conclusions

This study makes the first effort in examining the factors driving lightning activity in Alaska tundra with our statistical-dynamical modeling method. We demonstrated the effectiveness of integrating WRF and machine learning for lightning modeling in Alaskan tundra at 5 km spatial resolution. Our results provide insights into understanding the mechanisms of lightning-ignited fires in the tundra. We found PLI and weather variables at multiple pressure levels the most important predictors for modeling lightning potential in the tundra, indicating the primary role of convection in the formation of thunderstorms and CG lightning. Moreover, applicable to other data-scarce regions, our method can further support lightning and fire prediction as well as fire management efforts in the HNL.

Acknowledgments

This work was supported by the NASA Terrestrial Ecology program grant NNX15AT79A.

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Please wait… references are loading.