ArticlePDF Available

Extreme Fire Season in California: A Glimpse Into the Future?

Authors:
S1
OCTOBER 2015AMERICAN METEOROLOGICAL SOCIETY |
2. EXTREME FIRE SEASON IN CALIFORNIA: A GLIMPSE
INTO THE FUTURE?
Jin-Ho Yoon, S.-Y. Simon Wang, RobeRt R. gillieS, laWRence HippS,
ben KRavitz, and pHilip J. RaScH
Introduction. California has been under drought
conditions since 2012, and the drought worsened
considerably in the winter of 2013/14 (e.g., Wang et
al. 2014), which fueled an extreme fire season in 2014
(Hart et al. 2015). The early onset of the 2014 dry sea-
son (Supplemental Fig. S2.1) fueled an extraordinary
jump in wildfires. Between 1 January and 20 Septem-
ber, the California Department of Forestry and Fire
Protection reported thousands more fires than the
five-year average (www.fire.ca.gov). In early August,
a state of emergency was declared for a single wildfire
that had burned 32 000 acres (http://gov.ca.gov/news.
php?id=18645). This unusual fire season is expected
to continue well through 2015.
The connection between a warming climate and
lengthened fire seasons may seem intuitive, given
the general tendency toward a hot-and-dry climate
scenario and an earlier snowmelt (Westerling et al.
2006). However, what is not yet fully understood is
the extent to which the projected wetter climate in
California towards the latter part of the 21st century
(Neelin et al. 2013) could affect wildfire risk in the
future; this historica l drought and unusual fire season
also calls attention to possible impacts from human-
induced climate change.
Satellite merged data of burned area from the
fourth generation of the Global Fire Emissions Data-
base (GFED4; Giglio et al. 2013) was analyzed (online
supplemental material). Because the GFED4 product
may underestimate wildfire extent due to its limit in
the minimum detectable burned area and obscura-
tion by cloud cover, the Keetch–Byram Drought index
(KBDI; Janis et al. 2002; Keetch and Byram 1968),
routinely used by the United States Forest Service
for monitoring fire risk, was included as well. The
KBDI is computed with both the observational and
simulated daily precipitation and maximum surface
temperature. Obser vational dataset is from the North
American Land Data Assimilation phase 2 (NLDAS2;
Xia et al. 2012).
Fire extent of 2014. Figure 2.1 shows the annual
mean KBDI, the fractional area under extreme fire
risk (online supplemental material), and the burned
area averaged for entire California (Fig. 2.1a) and
northern California—north of 39°N (Fig. 2.1b). Both
the KBDI and the extreme fire risk exhibit a steady
increase over California since 1979 despite the rather
large interannual fluctuation. In terms of area burned
in GFED4, 2014 ranks the sixth largest in the entire
state and second in northern California; but in terms
of the KBDI and the extreme fire risk, 2014 ranks
first in both the entire state and northern California.
Also noteworthy is that the two largest burned areas
in northern California, over the 18-year record of
GFED4, occurred in 2012 and 2014. Spatially, the
area of higher fire risk in 2014, that is, a KBDI value
higher than 400, extends further north compared to
that of 2012 (Figs. 2.1e,f), consistent with the burned
area (Figs. 2.1c,d).
Attribution and projection. Wildfire simulations and
projections are general ly performed using stand-alone
vegetation models (e.g., Brown et al. 2004; Cook et
al. 2012; Luo et al. 2013; Scholze et al. 2006; Yue et al.
2013) driven by global climate model output. While
the advantage of using a stand-alone vegetation
model lies in its application to high spatial resolution
through downscaling, disadvantages include added
uncertainty produced from downscaling (e.g., Shukla
and Lettenmaier 2013; Yoon et al. 2012). In this study,
The fire season in northern California during 2014 was the second largest in terms of burned areas since
1996. An increase in fire risk in California is attributable to human-induced climate change.
AFFILIATIONS: Yoon, KR avitz, an d RaScH Atmospheric
Sciences and Global Change Division, Pacific Northwest National
Laboratory, Richland, Washington; Wang, gil lieS, a nd HippS
Utah Climate Center/Dept. Plants, Soils and Climate, Utah State
University, Logan, Utah
D O I: 10 . 117 5 / BA M S - D -1 5 - 0 0 114 . 1
A supplement to this article is available online (10.1175
/BAM S-D -15 -00114.1)
S2 OCTOBER 2015
|
we analyzed both the KBDI and wildfire probabilities
computed directly within the Community Earth
System Model version 1 (CESM1), which are primar-
ily driven by the dryness of the surface soil and the
availability of fuel load, that is, vegetation (Thonicke
et al. 2001). Although CESM1’s spatial resolution of
1-degree is relatively coarse, the model does simulate
well the climate drivers of fire, such as precipitation
and surface air temperature of California (Wang et al.
2014). Further, the CESM1 has produced 30 members
(online supplemental material) spanning historical
(1920–2005) and future periods (2006–80; based on
RCP8.5 scenario), together with a pre-industrial sim-
ulation of 1800 years. These model outputs provide
a unique opportunity for a detection and attribution
study conducted here to assess wildfire probabilities
under climate change.
Projections for California did show a steady in-
crease of the fire risk based solely upon the KBDI
(Fig. 2.2a) and are consistent with recent studies
(Dennison et al. 2014; Lin et al. 2014; Luo et al. 2013;
van Mantgem et al. 2013) that indicate increased oc-
currence of area burned and wildfire intensity and
duration over the western United States. The CESM1
projects only a slight increase in annual precipitation
accompanied with increasing surface warming after
1990 through 2070 (Supplemental Figs. S2.2b,c), con-
sistent with those produced by the Coupled Model
Intercomparison Project Phase 5 (CMIP5) ensembles
(Neelin et al. 2013). At face value, these simulations
of a slightly wetter climate from 1990 onward could
explain the cessation of the simulated fire probability
increase at the end of 20th century (Supplemental Fig.
S2.2c). However, the KBDI and the extreme fire risk
measures, computed here in terms of the fractional
area and the extreme fire danger days (Figs. 2.2b,c),
do show a steady and rapid increase from early 1990s
and 2000s.
To what extent can the change of the extreme fire
risk over California be attributed to global warming?
First, we need to understand how much fluctuation
is caused by natural climate variability alone (e.g.,
Kitzberger et al. 2007). Analyzing the 1800-year
pre-industrial simulation of the CESM1 by treating
the simulation as 18 member ensembles of 100-year
simulation, the pre-industrial simulations envelops
entirely both the KBDI and the extreme fire risk mea-
sures f luctuation for the period spanning 1920–80
Fig. 2.1. Annual mean of the KBDI in black, fraction of the area that are under the extreme fire risk in red, which
is defined as KBDI > 600, and burned area from GFED4 in orange averaged for (a) California and (b) Northern
California. Spatial distribution of burned area in hectare (ha) from GFED4 averaged for (c) 2014 and (d) 2012,
and corresponding the annual mean KBDI in (e) and (f).
S3
OCTOBER 2015AMERICAN METEOROLOGICAL SOCIETY |
(Fig. 2.2). Beginning in the 1990s—the later part of
the historical simulation—a clear separation emerges
between the extreme fire risks driven by the anthro-
pogenic climate forcing and that of natural climate
variability. However, 2014 occurred in a period of
rapidly increasing extreme fire risk. The pace of
increasing extreme fire risk according to simulation
has accelerated since the early 21st century and is
expected to surpass the range of natural climate vari-
ability. Observations show much faster
increases of the KBDI and extreme fire
risk measures (gray lines in Fig. 2.2).
The accelerated increase in the
KBDI and the extreme fire risk in
relation to the projected wetter cli-
mate in California is intriguing. To
increase extreme fire risk, two basic
situations need to be present: one is
abundant fuel load (i.e., surface veg-
etation coverage enhanced through
precipitation), and the other is the
occurrence of a hot-and-dry climate
regime or drought to dry the vegeta-
tion. A process called CO2 fertilization
(Donohue et al. 2013) tends to increase
vegetation activity simply through the
uptake of an increasing atmospheric
CO2. Under such a scenario along with
a wetter climate, vegetation growth
would increase and subsequently
supply sufficient fuel load. Though
population growth and associated
urban area change are accounted for
in the model, the CESM1 produced
fire probability does not account for
incidence of human-caused fire igni-
tion, which correlate with population
growth.
The extent to which man-made
global warming has increased the
risk or strength of the recent drought
in California has been an active area
of research. The severity of the 2014
fire season in California was ana lyzed
and its potential link to anthropogenic
warming was suggested (Diffenbaugh
et al. 2015; Wang et al. 2015, 2014)
despite presence of natural climate
variability (Wang and Schubert 2014).
However, it is important to point out
from this study that, the increase in
extreme fire risk is expected within
the coming decade to exceed that of natural variabil-
ity and this serves as a n indication that anthropogenic
climate warming will likely play a significant role in
influencing California’s fire season.
Conclusions. The 2014 fire season saw the second larg-
est burned area in northern California since 1997,
next only to 2012, and ranks the highest since 1979
in the case of extreme fire risk over the entire state.
Fig. 2.2. (a) Annual mean of the KBDI from the large ensemble
simulation of the CESM1, (b) fractional area (%) under the extreme
fire risk, and (c) the extreme fire danger (days year1) over
California. Red (blue) indicates the historical and RCP8.5 (pre-
industrial) runs. Gray lines indicate 50% of the 2nd order trend of
the KBDI and the extreme fire risk measures based on the NLDAS2.
To remove the climatological bias, starting points are adjusted to
be the same as the modeled ensemble mean of year 1979.
S4 OCTOBER 2015
|
Although both fire measures are based upon observa-
tions, these derived variables do exhibit uncertainty
(Giglio et al. 2013; Xia et al. 2012). The recent extreme
fire seasons have occurred in a time of drought. Some
measures of extreme fire risk are also expected to in-
crease in the future despite the overall lack of change
in the mean f ire probability and annual precipitation
simulated by climate models for the next 50 years. Our
result, based on the CESM1 outputs, indicates that
man-made global warming is likely one of the causes
that will exacerbate the areal extent and frequency
of extreme fire risk, though the inf luence of internal
climate variability on the 2014 and the future fire
season is difficult to ascertain. .
ACKNOWLEDGEMENTS. Research by Yoon,
Kravitz, and Rasch was supported by the Earth
System Modeling program in the Office of Science/
DOE and Wang, and Gillies by the WaterSMART
grant from the Bureau of Reclamation. Computation
was done at the National Energy Research Scientific
Computing Center and the Environmental Molecular
Sciences Laboratory at PNNL. CESM1 is supported by
the NSF and DOE. PNNL is operated for the Depart-
ment of Energy by Battelle Memorial Institute under
Contract DEAC05-76RLO1830.
Brown, T., B. Hall, and A. Westerling, 2004: e impact
of twenty-rst century climate change on wildland
re danger in the western United States: An applica-
tions perspective. Climatic Change, 62, 365–388.
Cook, B., N. Zeng, and J.-H. Yoon, 2012: Will Amazo-
nia dry out? Magnitude and causes of change from
IPCC climate model projections. Earth Interact., 16,
1-27, doi: 10.1175/2011EI39 8.1.
Dennison, P. E., S. C. Brewer, J. D. Arnold, and M. A.
Moritz, 2014: Large wildre trends in the western
United States, 1984–2011. Geophy. Res. Lett., 41,
2928–2933, doi:10.1002/2014GL059576.
Dienbaugh, N. S., D. L. Swain, and D. Touma, 2015:
Anthropogenic warming has increased drought risk
in California. Proc. Natl. Acad. Sci. USA, 11 2, 3931–
3936, doi:10.1073/pnas.1422385112.
Donohue, R. J., M. L. Roderick, T. R. McVicar, and G.
D. Farquhar, 2013: Impact of CO2 fertilization on
maximum foliage cover across the globe’s warm,
arid environments. Geophys. Res. Lett., 40, 3031–
3035, doi:10.1002/grl.50563.
Giglio, L., J. T. Randerson, and G. R. van der Werf,
2013: Analysis of daily, monthly, and annual burned
area using the fourth-generation global re emis-
sions database (GFED4). J. Geophys. Res. Biogeosci.,
118, 317–328, doi:10.1002/jgrg.20042.
Hart, S. J., T. Schoennagel, T. T. Veblen, and T. B. Chap-
man, 2015: Area burned in the western United States
is unaected by recent mountain pine beetle out-
breaks. Proc. Natl. Acad. Sci. USA, 112 , 4375–4380,
doi:10.1073/pnas.1424037112.
Janis, M. J., M. B. Johnson, and G. Fort hun, 2002: Near-
real time mapping of Keetch-Byram drought index
in the south-eastern United States. Int. J. Wildland
Fire, 11, 281–289.
Keetch, J. J., and G. M. Byram, 1968: A drought index
for forest re control. U.S.D.A. Forest Service Res.
Paper SE-38, 32 pp. [Available online at www.srs.
fs.usda.gov/pubs/rp/rp_se038.pdf.]
Kitzberger, T., P. M. Brown, E. K. Heyerdahl, T. W.
Swetnam, and T. T. Veblen, 2007: Contingent Pacif-
ic–Atlantic Ocean inuence on multicentury wild-
re synchrony over western North America. Proc.
Natl. Acad. Sci. USA, 104 , 543–548.
Lin, H.-W., J. L. McCarty, D. Wang, B. M. Rogers, D.
C. Morton, G. J. Collatz, Y. Jin, and J. T. Randerson,
2014: Management and climate contributions to
satellite-derived active re trends in the contiguous
United States. J. Geophys. Res. Biogeosci., 119, 645–
660, doi:10.1002/2013JG002382.
Luo, L., Y. Tang, S. Zhong, X. Bian, and W. E. Heilman,
2013: Will future climate favor more erratic wildres
in the western United States? J. Appl. Meteor. Clima-
tol., 52, 2410 –2417, doi :10.1175/JAMC-D-12- 0317.1.
Mo, K. C., and R. W. Higgins, 1998a: Tropical convec-
tion and precipitation regimes in the western United
States. J. Climate, 11, 2404–2423.
and
, 1998b: Tropical inuences on California
precipitation. J. Climate, 11, 412–430.
Neelin, J. D., B. Langenbrunner, J. E. Meyerson, A.
Hall, and N. Berg, 2013: California winter precipita-
tion change under global warming in the Coupled
Model Intercomparison Project Phase 5 Ensem-
ble. J. Climate, 26, 6238– 6256, doi:10.1175/JCLI-
D-12-00514.1.
Scholze, M., W. Knorr, N. W. Arnell, and I. C. Pren-
tice, 2006: A climate-change risk analysis for world
ecosystems. Proc. Natl. Acad. Sci. USA, 103, 13116–
1312 0.
REFERENCES
S5
OCTOBER 2015AMERICAN METEOROLOGICAL SOCIETY |
Shukla, S., and D. P. Lettenmaier, 2013: Multi-RCM
ensemble downscaling of NCEP CFS winter season
forecasts: Implications for seasonal hydrologic fore-
cast skill. J. Geophys. Res. Atmos., 118 , 10770–1079 0,
doi:10.1002/jgrd. 50628.
onicke, K., S. Venevsky, S. Sitch, and W. Cramer,
2001: e role of re disturbance for global vegeta-
tion dynamics: Coupling re into a Dynamic Global
Vegetation Model. Global Ecol. Biogeogr., 10, 661–
67 7.
van Mantgem, P. J., J. C. B. Nesmith, M. Keifer, E. E.
Knapp, A. Flint, and L. Flint, 2013: Climatic stress
increases forest re severity across the western
United States. Ecol. Lett., 16, 1151–1156, doi:10.1111/
ele.12151.
Wang, H., and S. Schubert, 2014: Causes of the extreme
dry conditions over California during early 2013 [in
“Explaining Extreme Events of 2013 from a Climate
Perspective”]. Bull. Amer. Meteor. Soc., 95 (9), S7–
S11.
Wang, S.-Y. S., W.-R. Huang, and J.-H. Yoon, 2015: e
North American winter ‘dipole’ and extremesactiv-
ity: A CMIP5 assessment. Atmos. Sci. Lett., 16, 338
345, doi:10.1002/asl2.565.
, L. Hipps, R . R. Gillies, and J.-H. Yoon, 2014: Prob-
able causes of the abnormal ridge accompanying the
2013-2014 California drought: ENSO precursor and
anthropogenic warming footprint. Geophys. Res.
Lett., 41, 3220–3226, doi:10.1002/2014GL059748.
Westerling, A. L., H. G. Hidalgo, D. R. Cayan, and T.
W. Swetnam, 2006: Warming and earlier spring in-
crease western US forest wildre activity. Science,
313, 940–943.
Xia, Y. L., and Coauthors, 2012: Continental-scale wa-
ter and energy ux analysis and validation for the
North American Land Data Assimilation System
project phase 2 (NLDAS-2): 1. Intercomparison and
application of model products. J. Geophys. Res., 117,
D03109, doi:10.1029/2011JD0160 48 .
Yoon, J. H., K. Mo, and E. F. Wood, 2012: Dynamic-
model-based seasonal prediction of meteorological
drought over the contiguous United States. J. Hydro-
mete or., 13, 463– 482, doi:10.1175/JH M-D-11-038.1.
Yue, X., L. J. Mickley, J. A. Logan, and J. O. Kaplan,
2013: Ensemble projections of wildre activ-
ity and carbonaceous aerosol concentrations over
the western United States in the mid-21st century.
Atmos. Environ., 77, 767–780, doi:10.1016/j.at-
mosenv.2013.06.003.
... A systematic review on the assessment of learning outcomes and human factors in the context of VR firefighting training found that the majority of studies chose an urban setting and focused on simulating search and rescue tasks (Bliss et al., 1997;Tate et al., 1997;Backlund et al., 2007;Jeon et al., 2019). Two studies differed in scope: Cohen-Hatton and Honey (2015), which evaluated the training of cognitive skills of commanders, and Clifford et al. (2020), which focused on aerial firefighting and simulating radio communication disruptions in a rural setting. In terms of domains of knowledge acquisition (Krathwohl, 2002), apart from Cohen-Hatton and Honey (2015), none of the reviewed studies focused on factual, conceptual, or metacognitive knowledge, instead focusing on the acquisition of procedural knowledge and the accurate replication of real-world tasks. ...
... Two studies differed in scope: Cohen-Hatton and Honey (2015), which evaluated the training of cognitive skills of commanders, and Clifford et al. (2020), which focused on aerial firefighting and simulating radio communication disruptions in a rural setting. In terms of domains of knowledge acquisition (Krathwohl, 2002), apart from Cohen-Hatton and Honey (2015), none of the reviewed studies focused on factual, conceptual, or metacognitive knowledge, instead focusing on the acquisition of procedural knowledge and the accurate replication of real-world tasks. Rural firefighting settings and factual, conceptual, or metacognitive knowledge acquisition remain largely underexplored. ...
... To posit suitable tasks to achieve our overall ILOs, we followed a similar approach as Fowler (2015) by listing the learning stages and associated learning outcomes followed by listing general learning tasks [based on the work of Conole et al. (2004)] that could implement the learning outcome (see Table 1). However, Conole et al. (2004) refer to general learning tasks as "mini-learning activities, " which we refrain from to maintain the distinction between learning tasks and learning activities as previously defined. ...
Article
The use of virtual reality (VR) in firefighter training is promising because it provides cost-effective, safe environments that arouse similar behavioral responses to real-life scenarios. However, the pedagogical potential of VR and its impact on learning outcomes compared to traditional methods is currently an under-explored area. This research investigates how well VR can support learning compared to traditional methods in the context of training firefighters in combating vegetation fires. A VR learning environment was developed, informed by a “design for learning” framework providing a pedagogical underpinning. A between-subjects experiment was conducted with 40 participants to measure the knowledge transfer of the VR learning environment against the official textbook. In addition, VR's theorized learning benefits of intrinsic motivation, situational interest, and self-efficacy were compared with textbook-based learning. Lastly, the design quality of the learning environment was assessed based on its learning and user experience. We employed a primarily quantitative approach to data collection and analysis, using a combination of knowledge test results and questionnaires, with supporting qualitative data from semi-structured interviews and observation notes to answer our hypotheses. The results found a significant difference between the knowledge transfer of both conditions, with textbook-based learning more effectively transferring factual and conceptual knowledge than VR. No significant difference was found in reported self-efficacy between the two conditions but was found in reported levels of intrinsic motivation and situational interest, which were higher in the VR condition. The design was found to have facilitated a good user and learning experience, assessed via questionnaire responses. During interviews, VR participants reported high levels of satisfaction with the experience, praising the hands-on learning approach and interactivity, while reporting frustration with the lack of knowledge reinforcement and initial difficulties with the controls. A key finding was that presence was found to be negatively associated with knowledge transfer, which we theorize to be caused by the novelty of the realistic VR environment distracting participants from the more familiar lesson content. This research contributes to the body of work related to knowledge transfer within VR in this domain while highlighting key pedagogical and design considerations that can be used to inform future design implementations.
... Therefore, it is not enough to only consider changes in re weather, and coupled re-vegetation models are required for a complete picture of change. Previous attribution studies focusing on vegetation res have often used re danger indices [10][11][12][13][14][15][16][17][18][19][20][21][22] , drought indices [22][23][24] , or individual components of re weather such as temperature and preciptaton 25,26 which misses changes in ignitions and fuel. Observation-based approaches to modelling re and its controls have proven useful for explaining how recent fuel and ignition changes in uence burning 3,27,28 . ...
... Therefore, it is not enough to only consider changes in re weather, and coupled re-vegetation models are required for a complete picture of change. Previous attribution studies focusing on vegetation res have often used re danger indices [10][11][12][13][14][15][16][17][18][19][20][21][22] , drought indices [22][23][24] , or individual components of re weather such as temperature and preciptaton 25,26 which misses changes in ignitions and fuel. Observation-based approaches to modelling re and its controls have proven useful for explaining how recent fuel and ignition changes in uence burning 3,27,28 . ...
... Therefore, it is not enough to only consider changes in re weather, and coupled re-vegetation models are required for a complete picture of change. Previous attribution studies focusing on vegetation res have often used re danger indices [10][11][12][13][14][15][16][17][18][19][20][21][22] , drought indices [22][23][24] , or individual components of re weather such as temperature and preciptaton 25,26 which misses changes in ignitions and fuel. Observation-based approaches to modelling re and its controls have proven useful for explaining how recent fuel and ignition changes in uence burning 3,27,28 . ...
... Therefore, it is not enough to only consider changes in re weather, and coupled re-vegetation models are required for a complete picture of change. Previous attribution studies focusing on vegetation res have often used re danger indices [10][11][12][13][14][15][16][17][18][19][20][21][22] , drought indices [22][23][24] , or individual components of re weather such as temperature and preciptaton 25,26 which misses changes in ignitions and fuel. Observation-based approaches to modelling re and its controls have proven useful for explaining how recent fuel and ignition changes in uence burning 3,27,28 . ...
Preprint
Full-text available
Fires are now raging longer and more intensely in many regions worldwide. However, non-linear interactions between fire weather, fuel, land use, management, and ignitions so far impeded formal attribution of global burned area changes. Here we show that climate change is increasingly explaining regional burned area patterns, using an ensemble of global fire models. Climate change has increased global burned area by 16% for the period 2003–2019, and raised the probability of experiencing months with above-average global burned area by 43%. Climate change-induced burned area increased in most regions, including those recently struck by extreme fire like Australia, South America and Siberia. Moreover, the contribution of climate change to burned area is rising by 0.22%year − 1 globally, with the largest increase in Central Australia (2.5%year − 1 ). Our results highlight the importance of immediate, drastic and sustained greenhouse gas emission reductions to stabilize fire impacts on lives, livelihoods and ecosystems.
... Wildfires pose serious threats to society, environment, and ecosystems as they can disrupt, damage, and destroy infrastructure, services, and properties (Barbero et al. 2015, Thomas et al. 2017. These threats are expected to grow with increasing severity and frequency of wildfires due to climate-pattern change, fire management, and the expansion of the wildland-urban interface (Abatzoglou & Williams 2016, Yoon et al. 2015. ...
Preprint
Full-text available
Background. Wildfire research uses ensemble methods to analyze fire behaviors and assess uncertainties. Nonetheless, current research methods are either confined to simple models or complex simulations with limits. Modern computing tools could allow for efficient, high-fidelity ensemble simulations. Aims. This study proposes a high-fidelity ensemble wildfire simulation framework for studying wildfire behavior, ML tasks, fire-risk assessment, and uncertainty analysis. Methods. In this research, we present a simulation framework that integrates the Swirl-Fire large-eddy simulation tool for wildfire predictions with the Vizier optimization platform for automated run-time management of ensemble simulations and large-scale batch processing. All simulations are executed on tensor-processing units to enhance computational efficiency. Key results. A dataset of 117 simulations is created, each with 1.35 billion mesh points. The simulations are compared to existing experimental data and show good agreement in terms of fire rate of spread. Computations are done for fire acceleration, mean rate of spread, and fireline intensity. Conclusions. Strong coupling between these 2 parameters are observed for the fire spread and intermittency. A critical Froude number that delineates fires from plume-driven to convection-driven is identified and confirmed with literature observations. Implications. The ensemble simulation framework is efficient in facilitating parametric wildfire studies.
... Recent fire seasons in California have been among the worst on record, part of a trend of larger and more severe wildfires driven in part by climate change [1][2][3][4]. Both the area burned by wildfire and the incidence of severe wildfires and megafires are projected to increase in the coming decades [2]; however, management practices have the potential to mitigate these projected changes substantially [2,5,6]. ...
Article
Full-text available
Between 2010 and 2020, an average of 36,037 hectares of grassland burned in wildfires in California each year, emitting greenhouse gasses (GHGs) and particulate matter (PM). These emissions impact climate and human health. Cattle grazing removes herbaceous fuel through the consumption of forage; however, ruminant digestion also emits GHGs. The purpose of this study was to examine the GHG and PM impact of livestock grazing in grasslands that go on to burn. We used Monte Carlo simulation to determine whether forage consumption by livestock led to reductions in grassland wildfire emissions and whether these reductions outweighed the emissions from the digestion of that forage. We estimate that between 2010 and 2020, an average of 11,590 metric tons (MT) of herbaceous fuel were removed by cattle annually from grasslands in California that went on to burn. This resulted in annual wildfire emission reductions ranging between 0.001 and 0.025 million metric tons (MMT) of CO2 equivalents (CO2e) and between 11 and 314 MT of PM2.5; a small fraction of total GHG and PM emissions from wildfires in California. We also evaluated the change in emissions if burned grasslands in California’s Central and North Coast regions—where removing grazing can lead to the encroachment of shrubs into grasslands—were instead shrublands. If the grasslands that burned in these regions in 2020 had instead been shrublands, we estimate that as much as 0.90 MMT more CO2e and 8448 MT more PM2.5 would have been emitted by wildfires, highlighting the long-term implications of livestock grazing.
... Fire history is critical for understanding the potential effects of changing climate and human land use on the forest landscape [1][2][3][4]. Under the current climate change scenario, south-central Chile (37 • -42 • S), characterized by sharp declines in rainfall and increases in the recurrence of drought and heat waves [5][6][7][8], is expected to see an increase in the recurrence and magnitude of forest fires [9][10][11][12][13][14][15][16]. Fire occurrence in this area has also been linked to the climatic effects of Southern Annular Mode (SAM) [17,18] as well as to El Niño Southern Oscillation (ENSO) variability [17,19,20]. ...
Article
Full-text available
Multiproxy reconstructions of fire regimes in forest ecosystems can provide a clearer understanding of past fire activity and circumvent some limitations of single proxy reconstructions. While inferring fire history from scars in trees is the most precise method to reconstruct temporal fire patterns, this method is limited in Araucaria araucana forests by rot after fire injuries, successive fires that destroy the evidence and the prohibition of sample extraction from living Araucaria trees. In this context, dendrochemical studies in Araucaria trees and charcoal analysis from sediment cores can complement and extend the time perspective of the fire history in the relictual Araucaria-Nothofagus forests of the coastal range. We used dendrochemical, fire scar and charcoal records from the Nahuelbuta Coastal Range (37.8° S; 73° W) spanning the last 1000 years to reconstruct the fire history. The results indicate that periods with higher fire activity occurred between 1400 and 1650 AD. Long-term changes in the fire regime are related to increased climate variability over the last 1000 years, and especially with the arrival of settlers to the area after 1860 CE. The most severe fire events in the Nothofagus and Araucaria forests occurred when suitable fire-prone conditions were superimposed with high human densities.
... Wildfires are important natural disturbances that provide valuable ecosystem services, such as enabling vegetation to regenerate, reducing flammable biomass, and controlling pests and have been used for centuries by humans to this end (Barnes and Van Lear 1998, Scharenbroch et al. 2012, Hankins 2013, Ryan et al. 2013, Pausas and Keeley 2019. However, in North America and particularly in the western U.S.A., wildfire regimes are beginning to occur at greater frequencies and intensities than previously recorded due to long-term fire suppression, drought, a warming climate, and other natural and anthropogenic disturbances (Dennison et al. 2014, Yoon et al. 2015. Pausas and Keeley (2019) proposed that the modern "anthropogenic wildfire regime" may accrue fewer benefits to ecosystems than did historical wildfires or prescribed burns. ...
Article
Despite increasing severity and frequency of wildfires, knowledge about how fire impacts the ecology of tick-borne pathogens is limited. In 2018, the River Fire burned a forest in the far-western U.S.A. where the ecology of tick-borne pathogens had been studied for decades. Forest structure, avifauna, large and small mammals, lizards, ticks, and tick-borne pathogens (Anaplasma phagocytophilum, Borrelia burgdorferi, Borrelia miyamotoi) were assessed after the wildfire in 2019 and 2020. Burning reduced canopy cover and eliminated the layer of thick leaf litter that hosted free-living ticks, which over time was replaced by forbs and grasses. Tick abundance and the vertebrate host community changed dramatically. Avian species adapted to cavity nesting became most prevalent, while the number of foliage-foraging species increased by 83% as vegetation regenerated. Nine mammalian species were observed on camera traps, including sentinel (black-tailed jackrabbits) and reservoir hosts (western gray squirrels) of B. burgdorferi. One Peromyscus sp. mouse was captured in 2019 but by 2020, numbers were rebounding (n=37), although tick infestations on rodents remained sparse (0.2/rodent). However, western fence lizards (n=19) hosted 8.6 ticks on average in 2020. Assays for pathogens found no B. miyamotoi in either questing or host-feeding ticks, A. phagocytophilum DNA in 4% (1/23) in 2019, and 17% (29/173) in 2020 for questing and host-feeding ticks combined, and B. burgdorferi DNA in just 1% of all ticks collected in 2020 (2/173). We conclude that a moderately severe wildfire can have dramatic impacts on the ecology of tick-borne pathogens, with changes posited to continue for multiple years.
Chapter
A forest fire is an unplanned and unexpected event that contaminates our environment by burning the grass, trees, and vegetables of any forest and have an impact on the country’s economic model and the functioning of our ecosystem. A forest fire begins as a result of a natural occurrence, human activities, global warming, or lightning strikes. California is one of them that is affected by fires everyday. However, the statistics and procedures needed to reliably characterize fire propagation, behavior, and consequences are still lacking. As a result, our understanding of wildfire behavior remains restricted. In this paper, we have analyzed the area affected by the fire of August 7, 2016, with the help of machine intelligence deep learning model from Landsat 8 OLI images. The model has a total of five hidden layers, and the model has been trained with seven input features. The accuracy of the model has been obtained as 99.4712% with a baseline error of 0.53%. The area affected by a large fire to the north of San Bernardino in the state of California (USA) is 171.730934 km2 (42,435.637956 Acres). We used this approach to map California’s fire history in August 2016.
Article
Full-text available
Plain Language Summary Wildland‐urban interface (WUI) fires are increasing in the United States and around the world as the built environment continues to expand into the wildland. To better inform real‐time management of active wildfires, it is critical that the scientific community can better predict WUI fire spread. In this study, we rely on multiple observational platforms, including the “Doppler on Wheels” radar, to investigate the performance of a state‐of‐the‐art fire behavior model that links with a weather model during the Marshall Fire, which was a recent WUI fire that occurred in Colorado. While the modeling system performs well during the fire's initial propagation in fine fuels, it is unable to accurately predict spread in the built environment. While turbulence‐resolving simulations can accurately represent atmospheric flow features, more reliable predictability of wildfire behavior in the WUI will require consideration of urban fuels and fire ember spotting.
Article
Drought is a natural phenomenon that is mainly caused by less than normal precipitation over a long period that occurs in almost all climates. Since the late 1970s, the occurrence of drought has increased globally due to increased evaporation (caused by global warming). This phenomenon occurs slowly and can be short-term or terminated within a few months or continue for several years. drought has increased significantly in Iran in recent decades. The increase in droughts has many negative consequences in different areas of agriculture, water resources, and wildfire. The purpose of this study is to investigate the drought situation in the vegetation areas of Iran using the Keetch-Byram drought index (KBDI). For this purpose, the KBDI drought index derived from the data set of the European Centre for Medium-Range Weather Forecasts (ECMWF) fifth edition (ERA5) with a horizontal resolution of 0.25o during the period of 1981-2020 has been used. Also, the combined product of surface soil moisture (SSM) from microwave sensors AMI-WS (ERS-1 and ERS-2 satellites) and ASCAT (MetOpA-B satellite) with a horizontal resolution of 0.25 o to evaluate the KBDI drought index results as well as the moisture condition. The results showed that the khalij-Omani and Irani-Turani vegetation areas have very low soil moisture and high drought in Iran. The amount of soil moisture in these areas is low and the area-averaged is equal to 12.65% and 18.42%, respectively, which has been somewhat constant in all months, in other words, dryness is the predominant climate feature of these areas. The monthly analysis of the KBDI index shows the spread of drought severity towards the west, i.e., the vegetation area of Zagros, in the hot months of the year. A decrease in soil moisture and dryness can be observed in a major part of Iran from early spring to mid-autumn. Also, in the Zagros region, from mid-summer to mid-autumn, increasing values of the drought index and decreasing soil moisture can be observed. The minimum drought index is in the vegetation areas of Hyrkani and Arsbaran. The correspondence between the results of the surface soil moisture (SSM) of the satellite product and the KBDI drought index obtained from ECMWF-ERA5 shows the appropriate efficiency of the KBDI drought index in identifying drought centers in Iran and investigating the spatio-temporal patterns of drought occurrence. frequent low-intensity drought may favor more drought-tolerant species and adapt forests and grasslands to future conditions without the need for management measures.
Article
Full-text available
Significance Across western North America, abundant susceptible pine hosts and a suitable climate during the early 21st century have promoted widespread mountain pine beetle (MPB) outbreaks, leading to concern that dead fuels may increase wildfire risk. The assumption that outbreaks raise fire risk is driving far-reaching policy decisions involving expenditures of hundreds of millions of dollars. Contrary to the expectation that an MPB outbreak increases fire risk, spatial overlay analysis shows no effect of outbreaks on subsequent area burned during years of extreme burning across the West. These results refute the assumption that increased bark beetle activity has increased area burned; therefore, policy discussions should focus on societal adaptation to the effects of the underlying drivers: warmer temperatures and increased drought.
Article
Full-text available
Wildfires that occurred over the western United States during August 2012 were fewer in number but larger in size when compared with all other Augusts in the twenty-first century. This unique characteristic, along with the tremendous property damage and potential loss of life that occur with large wildfires with erratic behavior, raised the question of whether future climate will favor rapid wildfire growth so that similar wildfire activity may become more frequent as climate changes. This study addresses this question by examining differences in the climatological distribution of the Haines index (HI) between the current and projected future climate over the western United States. The HI, ranging from 2 to 6, was designed to characterize dry, unstable air in the lower atmosphere that may contribute to erratic or extreme fire behavior. A shift in HI distribution from low values (2 and 3) to higher values (5 and 6) would indicate an increased risk for rapid wildfire growth and spread. Distributions of Haines index are calculated from simulations of current (1971-2000) and future (2041-70) climate using multiple regional climate models in the North American Regional Climate Change Assessment Program. Despite some differences among the projections, the simulations indicate that there may be not only more days but also more consecutive days with HI 5 during August in the future. This result suggests that future atmospheric environments will be more conducive to erratic wildfires in the mountainous regions of the western United States.
Article
Full-text available
Significance California ranks first in the United States in population, economic activity, and agricultural value. The state is currently experiencing a record-setting drought, which has led to acute water shortages, groundwater overdraft, critically low streamflow, and enhanced wildfire risk. Our analyses show that California has historically been more likely to experience drought if precipitation deficits co-occur with warm conditions and that such confluences have increased in recent decades, leading to increases in the fraction of low-precipitation years that yield drought. In addition, we find that human emissions have increased the probability that low-precipitation years are also warm, suggesting that anthropogenic warming is increasing the probability of the co-occurring warm–dry conditions that have created the current California drought.
Article
Full-text available
The 2013-2014 winter in North America brought intense drought in the West and severe cold in the East. The circulation anomalies were characterized as a dipole: an amplified upper-level ridge over the West Coast and a deepened trough over the central-eastern U.S. A previous study using a single model has linked the dipole to the El Niño precursor and found that this link has strengthened in recent years. Here, 17 models from the Coupled Model Intercomparison Project Phase 5 are utilized to examine the dipole activity. Most models capture the dipole and its association with El Niño precursor and project this association to strengthen.
Article
Full-text available
The 2013-14 California drought was accompanied by an anomalous high-amplitude ridge system. The anomalous ridge was investigated using reanalysis data and the Community Earth System Model (CESM). It was found that the ridge emerged from continual sources of Rossby wave energy in the western North Pacific starting in late summer, and subsequently intensified into winter. The ridge generated a surge of wave energy downwind and deepened further the trough over the northeast U.S., forming a dipole. The dipole and associated circulation pattern is not linked directly with either ENSO or Pacific Decadal Oscillation; instead it is correlated with a type of ENSO precursor. The connection between the dipole and ENSO precursor has become stronger since the 1970s, and this is attributed to increased GHG loading as simulated by the CESM. Therefore, there is a traceable anthropogenic warming footprint in the enormous intensity of the anomalous ridge during winter 2013-14, the associated drought and its intensity.
Article
Projections of possible precipitation change in California under global warming have been subject to considerable uncertainty because California lies between the region anticipated to undergo increases in precipitation at mid-to-high latitudes and regions of anticipated decrease in the subtropics. Evaluation of the large-scale model experiments for phase 5 of the Coupled Model Intercomparison Project (CMIP5) suggests a greater degree of agreement on the sign of the winter (December-February) precipitation change than in the previous such intercomparison, indicating a greater portion of California falling within the increased precipitation zone. While the resolution of global models should not be relied on for accurate depiction of topographic rainfall distribution within California, the precipitation changes depend substantially on large-scale shifts in the storm tracks arriving at the coast. Significant precipitation increases in the region arriving at the California coast are associated with an eastward extension of the region of strong Pacific jet stream, which appears to be a robust feature of the large-scale simulated changes. This suggests that effects of this jet extension in steering storm tracks toward the California coast constitute an important factor that should be assessed for impacts on incoming storm properties for high-resolution regional model assessments.
Article
The 2013 SST anomalies produced a predilection for California drought, whereas the long-term warming trend appears to make no appreciable contribution because of the counteraction between its dynamical and thermodynamic effects.
Article
We assess the value of dynamical versus statistical downscaling of National Centers for Environmental Prediction's (NCEP) Climate Forecast System (CFS) winter season forecasts for seasonal hydrologic forecasting. Dynamically downscaled CFS forecasts for 1 December to 30 April of 1982–2003 were obtained from the Multi‐RCM Ensemble Downscaling (MRED) project that used multiple Regional Climate Models (RCMs) to downscale CFS forecasts. Statistical downscaling of CFS forecasts was achieved by a much simpler bias correction and spatial downscaling method. We evaluate forecast accuracy of runoff (RO), soil moisture (SM), and snow water equivalent produced by a hydrology model forced with dynamically (the MRED forecasts) and statistically downscaled CFS forecasts in comparison with predictions of those variables produced by forcing the same hydrology model with gridded observations (reference data set). Our results show that the MRED forecasts produce modest skill beyond what results from statistical downscaling of CFS. Although the improvement in hydrologic forecast skill associated with the ensemble average of the MRED forecasts (Multimodel) relative to statistical downscaled CFS forecasts is field significant for RO and SM forecasts with up to 3 months lead, the region of improvement is mainly limited to parts of the northwest and north central U.S. In general, one or more RCMs outperform the other RCMs as well as the Multimodel. Hence, we argue that careful selection of RCMs (based on their hindcast skill over any given region) is critical to improving hydrologic forecast skill using dynamical downscaling.
Article
[1] Satellite observations reveal a greening of the globe over recent decades. The role in this greening of the “CO2 fertilization” effect—the enhancement of photosynthesis due to rising CO2 levels—is yet to be established. The direct CO2 effect on vegetation should be most clearly expressed in warm, arid environments where water is the dominant limit to vegetation growth. Using gas exchange theory, we predict that the 14% increase in atmospheric CO2 (1982–2010) led to a 5 to 10% increase in green foliage cover in warm, arid environments. Satellite observations, analyzed to remove the effect of variations in precipitation, show that cover across these environments has increased by 11%. Our results confirm that the anticipated CO2 fertilization effect is occurring alongside ongoing anthropogenic perturbations to the carbon cycle and that the fertilization effect is now a significant land surface process.