Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 23 August 2024

Global patterns and drivers of post-fire vegetation productivity recovery

  • Hongtao Xu 1 ,
  • Hans W. Chen   ORCID: orcid.org/0000-0002-8601-6024 2 ,
  • Deliang Chen   ORCID: orcid.org/0000-0003-0288-5618 3 ,
  • Yingping Wang 4 ,
  • Xu Yue   ORCID: orcid.org/0000-0002-8861-8192 5 ,
  • Bin He   ORCID: orcid.org/0000-0002-9088-262X 1 , 6 ,
  • Lanlan Guo 1 ,
  • Wenping Yuan 7 ,
  • Ziqian Zhong   ORCID: orcid.org/0000-0003-2608-0953 2 ,
  • Ling Huang   ORCID: orcid.org/0000-0002-3262-151X 8 ,
  • Fei Zheng   ORCID: orcid.org/0000-0002-6897-1626 9 ,
  • Tiewei Li 1 &
  • Xiangqi He 1  

Nature Geoscience ( 2024 ) Cite this article

493 Accesses

10 Altmetric

Metrics details

  • Climate-change ecology
  • Fire ecology

Wildfires cause critical shifts in ecosystem functions, such as dramatic reductions in vegetation productivity. However, how fast vegetation regains its pre-fire productivity levels and the key influencing factors remain poorly understood on a global scale. Here we present the global estimates of post-fire vegetation productivity recovery from 2004 to 2021 using gross primary productivity observations and related proxies at a spatial resolution of 10 km, employing a random forest model to identify the key factors influencing recovery time. Roughly 87% of burned vegetation regained pre-fire productivity levels within 2 years, with evergreen needleleaf forests and savannas displaying the lengthiest recovery periods. During the recovery phase, post-fire climate conditions, such as soil moisture, vapour pressure deficit and air temperature, had nonlinear impacts on recovery time globally. These climatic factors exhibited a dominant role in regional recovery time in ~89% of the globally assessed area. As climate aridity decreased, the areas where recovery time was dominated by soil moisture and vapour pressure deficit decreased, while the influence of temperature increased. Soil-moisture-dominated regions witnessed reduced proportions of promoting vegetation recovery as aridity decreased, whereas vapour pressure deficit and air-temperature-dominated regions saw an increase in such proportions. Regions with strong human interventions were associated with accelerated vegetation recovery compared with similar ecosystems with smaller human interventions. These findings had important implications for global carbon-cycle assessments and fire-management strategies.

This is a preview of subscription content, access via your institution

Access options

Access Nature and 54 other Nature Portfolio journals

Get Nature+, our best-value online-access subscription

$29.99 / 30 days

cancel any time

Subscribe to this journal

Receive 12 print issues and online access

$259.00 per year

only $21.58 per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

research paper on fire ecology

Similar content being viewed by others

research paper on fire ecology

Fire may prevent future Amazon forest recovery after large-scale deforestation

research paper on fire ecology

Changes in land use enhance the sensitivity of tropical ecosystems to fire-climate extremes

research paper on fire ecology

Vegetation fires in the Anthropocene

Data availability.

All data used in this study are freely and publicly accessible. The source data underlying Figs. 1 – 4 and Extended Data Figs. 1 – 9 have been deposited in the Zenodo repository at https://zenodo.org/records/12669733 (ref. 67 ). The MCD64A1, MOD13Q1, ERA5-Land and GMTED datasets are derived from https://code.earthengine.google.com/ , the grid GPP products are available at https://globalecology.unh.edu/data/GOSIF-GPP.html , the ESA CCI land-cover products are available at https://www.esa-landcover-cci.org/?q=node/164 , the plant biodiversity data are available at https://ecotope.org/anthromes/biodiversity/plants/data/ , the soil sand content data are derived from the Regridded Harmonized World Soil Database v1.2 (ornl.gov) available at https://daac.ornl.gov/SOILS/guides/HWSD.html , the aridity index data are available at https://www.nature.com/articles/s41597-022-01493-1#data-availability and the annual human footprint datasets are available at https://figshare.com/articles/figure/An_annual_global_terrestrial_Human_Footprint_dataset_from_2000_to_2018/16571064 (ref. 68 ).

Code availability

The code for the analysis and mapping can be obtained from the Zenodo repository ( https://zenodo.org/records/12669733 ) 67 .

Zheng, B. et al. Record-high CO 2 emissions from boreal fires in 2021. Science 379 , 912–915 (2023).

Article   CAS   Google Scholar  

Li, F., Bond-Lamberty, B. & Levis, S. Quantifying the role of fire in the Earth system—part 2: impact on the net carbon balance of global terrestrial ecosystems for the 20th century. Biogeosciences 11 , 1345–1360 (2014).

Article   Google Scholar  

Tangney, R. et al. Success of post-fire plant recovery strategies varies with shifting fire seasonality. Commun. Earth Environ. 3 , 126 (2022).

Viana-Soto, A., García, M., Aguado, I. & Javier, S. Assessing post-fire forest structure recovery by combining LiDAR data and Landsat time series in Mediterranean pine forests. Int. J. Appl. Earth Obs. Geoinf. 108 , 102754 (2022).

Google Scholar  

Yue, X. & Unger, N. Fire air pollution reduces global terrestrial productivity. Nat. Commun. 9 , 5413 (2018).

Grünig, M., Seidl, R. & Senf, C. Increasing aridity causes larger and more severe forest fires across Europe. Glob. Change Biol. 29 , 1648–1659 (2023).

IPBES. Global assessment report on biodiversity and ecosystem services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services (version 1) (eds Brondízio, E. S. et al.). Zenodo https://doi.org/10.5281/zenodo.3831673 (2019).

Bousquet, E., Mialon, A., Rodriguez-Fernandez, N., Mermoz, S. & Yann, K. Monitoring post-fire recovery of various vegetation biomes using multi-wavelength satellite remote sensing. Biogeosciences 19 , 3317–3336 (2022).

Smith, T. & Boers, N. Reliability of vegetation resilience estimates depends on biomass density. Nat. Ecol. Evol. 7 , 1799–1808 (2023).

Jolly, W. M. et al. Climate-induced variations in global wildfire danger from 1979 to 2013. Nat. Commun. 6 , 7537 (2015).

Pausas, J. G. & Keeley, J. E. A burning story: the role of fire in the history of life. Bioscience 59 , 593–601 (2009).

Pausas, J. G. & Keeley, J. E. Wildfires and global change. Front. Ecol. Environ. 19 , 387–395 (2021).

Jae-Hyun, R. et al. Satellite-based evaluation of the post-fire recovery process from the worst forest fire case in South Korea. Remote Sens. 10 , 918 (2018).

Pérez-Cabello, F., Montorio, R. & Alves, D. B. Remote sensing techniques to assess post-fire vegetation recovery. Curr. Opin. Environ. Sci. Health 21 , 100251 (2021).

Shvetsov, E. G., Kukavskaya, E. A., Buryak, L. V. & Barrett, K. Assessment of post-fire vegetation recovery in Southern Siberia using remote sensing observations. Environ. Res. Lett. 14 , 055001 (2019).

Vianasoto, A., Aguado, I. & Salas, J. Identifying post-fire recovery trajectories and driving factors using landsat time series in fire-prone Mediterranean pine forests. Remote Sens. 12 , 1499 (2020).

Camila, V. J. et al. Drought-induced Amazonian wildfires instigate a decadal-scale disruption of forest carbon dynamics. Philos. Trans. R. Soc. Lond. B Biol. Sci. 373 , 20180043 (2018).

Chu, T. & Guo, X. Remote sensing techniques in monitoring post-fire effects and patterns of forest recovery in boreal forest regions: a review. Remote Sens. 6 , 470–520 (2013).

Faria, B. L. D., Marano, G., Piponiot, C., Silva, C. A. & Collalti, A. Model-BASED estimation of Amazonian forests recovery time after drought and fire events. Forests 12 , 8 (2020).

Heath, J. T., Chafer, C. J., Bishop, T. F. A. & Van Ogtrop, F. F. Post-fire recovery of eucalypt-dominated vegetation communities in the Sydney Basin, Australia. Fire Ecol. 12 , 53–79 (2016).

Weber, M. G. & Stocks, B. J. Forest fires and sustainability in the boreal forests of Canada. AMBIO 27 , 545–550 (1998).

Zhang, Q., Homayouni, S., Zhao, P. & Zhou, M. Burned vegetation recovery trajectory and its driving factors using satellite remote-sensing datasets in the Great Xing’An forest region of Inner Mongolia. Int. J. Wildland Fire 32 , 244–261 (2023).

Zheng, B. et al. Increasing forest fire emissions despite the decline in global burned area. Sci. Adv. 7 , eabh2646 (2021).

Zhang, P. et al. Using enhanced vegetation index and land surface temperature to reconstruct the solar-induced chlorophyll fluorescence of forests and grasslands across latitude and phenology. Front. Glob. Change 6 , 1257287 (2023).

Steiner, J. L., Wetter, J., Robertson, S., Teet, S. & Xiao, X. Grassland wildfires in the Southern Great Plains: monitoring ecological impacts and recovery. Remote Sens. 12 , 619 (2020).

Wagle, P. et al. Sensitivity of vegetation indices and gross primary production of tallgrass prairie to severe drought. Remote Sens. Environ. 152 , 1–14 (2014).

Huete, A. A comparison of vegetation indices over a global set of TM images for EOS-MODIS. Remote Sens. Environ. 59 , 440–451 (1997).

Gustau, C. V. et al. A unified vegetation index for quantifying the terrestrial biosphere. Sci. Adv. 7 , eabc7447 (2021).

Mutanga, O., Masenyama, A. & Sibanda, M. Spectral saturation in the remote sensing of high-density vegetation traits: a systematic review of progress, challenges and prospects. ISPRS-J. Photogramm. Remote Sens. 198 , 297–309 (2023).

Wu, S. et al. Global patterns and influencing factors of post-fire land cover change. Glob. Planet. Change 223 , 104076 (2023).

Li, X. & Xiao, J. Global climatic controls on interannual variability of ecosystem productivity: similarities and differences inferred from solar-induced chlorophyll fluorescence and enhanced vegetation index. Agric. Meteorol. 15 , 288–298 (2020).

Zhong, Z. et al. Disentangling the effects of vapor pressure deficit on northern terrestrial vegetation productivity. Sci. Adv. 9 , eadf316 (2023).

Liu, L. et al. Soil moisture dominates dryness stress on ecosystem production globally. Nat. Commun. 11 , 4892 (2020).

Chen, N. et al. Warming-induced vapor pressure deficit suppression of vegetation growth diminished in northern peatlands. Nat. Commun. 14 , 7885 (2023).

Tang, T., Shindell, D., Zhang, Y., Voulgarakis, A. & Samset, B. H. Response of surface shortwave cloud radiative effect to greenhouse gases and aerosols and its impact on summer maximum temperature. Atmos. Chem. Phys. 20 , 8251–8266 (2020).

Bicalho, E. et al. Temperature and light requirements for germination of species of Velloziaceae from different Brazilian rocky outcrops. Acta Bot. Bras. 32 , 240–246 (2018).

Choat, B. et al. Global convergence in the vulnerability of forests to drought. Nature 491 , 752–755 (2012).

Crawford et al. High temperature exposure increases plant cooling capacity. Curr. Biol. 22 , R396–R397 (2012).

Josef, U., Ingwers, M. W., Anne, M. G. M. & Teskey, R. O. Increase in leaf temperature opens stomata and decouples net photosynthesis from stomatal conductance in Pinus taeda and Populus deltoides x nigra . J. Exp. Bot. 68 , 1757–1767 (2017).

Lian, X. et al. Summer soil drying exacerbated by earlier spring greening of northern vegetation. Sci. Adv. 6 , eaax0255 (2020).

Dave, V. W., Guido, R., James, T. R., Brendan, M. R. & Yang, C. Global biomass burning fuel consumption and emissions at 500-m spatial resolution based on the Global Fire Emissions Database (GFED). Geosci. Model Dev. 15 , 8411–8437 (2022).

Sitch, S. et al. Evaluation of ecosystem dynamics, plant geography and terrestrial carbon cycling in the LPJ dynamic global vegetation model. Glob. Change Biol. 9 , 161–185 (2003).

Thonicke, K., Venevsky, S., Sitch, S. & Cramer, W. The role of fire disturbance for global vegetation dynamics: coupling fire into a dynamic global vegetation model. Glob. Ecol. Biogeogr. 10 , 661–677 (2001).

Bond, W. J. & Keane, R. E. Fires, ecological effects of. Reference Module in Life Sciences https://doi.org/10.1016/B978-0-12-809633-8.02098-7 (2017).

Johnson, D. B., Woolet, J., Yedinak, K. M. & Whitman, T. Experimentally determined traits shape bacterial community composition one and five years following wildfire. Nat. Ecol. Evol. 7 , 1419–1431 (2023).

Pellegrini, A. F. A. et al. Soil carbon storage capacity of drylands under altered fire regimes. Nat. Clim. Change 13 , 1089–1094 (2023).

IPCC. Summary for policymakers. In Climate Change 2021 : The Physical Science Basis (eds Masson-Delmotte, V. et al.) (Cambridge Univ. Press, 2021).

Giglio, L., Justice, C., Boschetti, L. & Roy, D. MCD64A1 MODIS/Terra+Aqua Burned Area Monthly L3 Global 500 m SIN Grid V006 [dataset]. NASA EOSDIS Land Process. Distrib. Act. Arch. Cent. https://doi.org/10.5067/MODIS/MCD64A1.006 (2015).

Qin, Y. et al. Improved estimates of forest cover and loss in the Brazilian Amazon in 2000–2017. Nat. Sustain. 2 , 764–772 (2019).

Li, X. & Xiao, J. Mapping photosynthesis solely from solar-induced chlorophyll fluorescence: a global, fine-resolution dataset of gross primary production derived from OCO-2. Remote Sens. 11 , 2563 (2019).

Gensheimer, J., Turner, A. J., Khler, P., Frankenberg, C. & Chen, J. A convolutional neural network for spatial downscaling of satellite-based solar-induced chlorophyll fluorescence (SIFnet). Biogeosciences 19 , 1777–1793 (2021).

Wang, Q., Moreno-Martínez, Á., Muñoz-Marí, J., Campos-Taberner, M. & Camps-Valls, G. Estimation of vegetation traits with kernel NDVI. ISPRS-J. Photogramm. Remote Sens. 195 , 408–417 (2023).

Zhang, Z. & Li, X. The resilience of ecosystems to drought. Glob. Change Biol. 29 , 3517–3518 (2023).

Didan, K. MOD13Q1 MODIS/Terra Vegetation Indices 16-Day L3 Global 250 m SIN Grid V006. NASA EOSDIS Land Process. DAAC https://doi.org/10.5067/MODIS/MOD13Q1.006 (2015).

Yao, Y., Liu, Y., Zhou, S., Song, J. & Fu, B. Soil moisture determines the recovery time of ecosystems from drought. Glob. Change Biol. 29 , 3562–3574 (2023).

Nemani et al. Climate-driven increases in global terrestrial net primary production from 1982 to 1999. Science 300 , 1560–1563 (2003).

Chen, Z., Wang, W., Forzieri, G. & Cescatti, A. Transition from positive to negative indirect CO 2 effects on the vegetation carbon uptake. Nat. Commun. 15 , 1500 (2024).

Land Cover CCI Product User Guide Version 2 Technical Report (ESA, 2017); maps.elie.ucl.ac.be/CCI/viewer/download/ESACCI-LC-Ph2-PUGv2_2.0.pdf

Oliver, C. D. & Larson, B. A. Forest Stand Dynamics (Wiley, 1996).

Nolan, R. H. et al. Limits to post‐fire vegetation recovery under climate change. Plant. Cell Environ. 44 , 3471–3489 (2021).

Ellis, E. C., Antill, E. C. & Kreft, H. All is not loss: plant biodiversity in the anthropocene. PLoS ONE 7 , e30535 (2012).

Huang et al. Accelerated dryland expansion under climate change. Nat. Clim. Change 6 , 166–171 (2016).

Wieder, W. R., Boehnert, J., Bonan, G. B. & Langseth, M. Regridded Harmonized World Soil Database v1.2 data set. Oak Ridge National Laboratory Distributed Active Archive Center https://daac.ornl.gov/SOILS/guides/HWSD.html (2014).

Danielson, J. J. & Gesch, D. B. Global Multi-resolution Terrain Elevation Data 2010 (GMTED2010) (USGS numbered series no. 2011–1073) (US Geological Survey, 2011).

Mu, H. et al. A global record of annual terrestrial Human Footprint dataset from 2000 to 2018. Sci. Data 176 , 176 (2022).

Chen, Y., Cheng, X., Liu, A., Chen, Q. & Wang, C. Tracking lake drainage events and drained lake basin vegetation dynamics across the Arctic. Nat. Commun. 14 , 73 (2023).

He, B. & Xu, H. Global patterns and drivers of post-fire vegetation productivity recovery [dataset]. Zenodo https://zenodo.org/records/12669733 (2024).

Mu, H. et al. An annual global terrestrial Human Footprint dataset from 2000 to 2018. figshare https://doi.org/10.6084/m9.figshare.16571064.v7 (2021).

Download references

Acknowledgements

This work was funded by the National Natural Science Foundation of China (grant no. 42361144877), State Key Laboratory of Earth Surface Processes and Resource Ecology (grant no. 2023-KF-02) and BNU-FGS Global Environmental Change Program (grant no. 2023-GC-ZYTS-01). H.W.C. and D.C. were supported by the Swedish national strategic research programs BECC and MERGE.

Author information

Authors and affiliations.

State Key Laboratory of Earth Surface Processes and Resource Ecology, Faculty of Geographical Science, Beijing Normal University, Beijing, China

Hongtao Xu, Bin He, Lanlan Guo, Tiewei Li & Xiangqi He

Department of Space, Earth and Environment, Division of Geoscience and Remote Sensing, Chalmers University of Technology, Gothenburg, Sweden

Hans W. Chen & Ziqian Zhong

Department of Earth Sciences, University of Gothenburg, Gothenburg, Sweden

Deliang Chen

CSIRO Oceans and Atmosphere, Aspendale, Victoria, Australia

Yingping Wang

Jiangsu Key Laboratory of Atmospheric Environment Monitoring and Pollution Control, Collaborative Innovation Center of Atmospheric Environment and Equipment Technology, School of Environmental Science and Engineering, Nanjing University of Information Science & Technology, Nanjing, China

Akesu National Station of Observation and Research for Oasis Agro-ecosystem, Xinjiang, China

Institute of Carbon Neutrality, Sino-French Institute for Earth System Science College of Urban and Environmental Sciences, Peking University, Beijing, China

Wenping Yuan

College of Urban and Environmental Sciences, Peking University, Beijing, China

International Center for Climate and Environment Science, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing, China

You can also search for this author in PubMed   Google Scholar

Contributions

B.H., L.G. and H.X. designed the research; H.X. performed the analysis and wrote the paper. H.W.C., D.C., Y.W. and X.Y. provided comments to improve the paper. B.H. supervised the project. H.W.C., L.G., W.Y., Z.Z., L.H., F.Z., T.L. and X.H. offered thoughts on the analysis and contributed to the writing of the paper.

Corresponding author

Correspondence to Bin He .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Peer review

Peer review information.

Nature Geoscience thanks Helen Poulos, Ryan Tangney and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Xujia Jiang, in collaboration with the Nature Geoscience team.

Additional information

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

Extended data

Extended data fig. 1 average recovery time globally and in arid, semi-arid, semi-humid, and humid regions..

a–c , The recovery time estimated by GPP ( a ), kNDVI ( b ), and EVI ( c ), respectively. Boxplots show the statistics – minima: first interquartile subtract 1.5 times the interquartile range, maxima: third interquartile add 1.5 times the interquartile range, horizontal lines: median; dots: mean; boxes: interquartile range; whiskers: 1.5 times the interquartile range. The interquartile range denote the difference between the third and first interquartile. The numbers in parentheses represent the sample sizes.

Extended Data Fig. 2 Average recovery time in the Southern and Northern Hemispheres estimated by GPP, kNDVI, and EVI, respectively.

Boxplots show the statistics – minima: first interquartile subtract 1.5 times the interquartile range, maxima: third interquartile add 1.5 times the interquartile range, horizontal lines: median; dots: mean; boxes: interquartile range; whiskers: 1.5 times the interquartile range. The interquartile range denote the difference between the third and first interquartile. The numbers in parentheses represent the sample sizes.

Extended Data Fig. 3 Average recovery time in six continents (Africa, Asia, Europe, North America, Oceania, and South America).

Extended data fig. 4 average recovery time in different vegetation types across continents..

a–g , Recovery time comparisons across different continents in evergreen broadleaf forest ( a ), deciduous broadleaf forest ( b ), evergreen needleleaf forest ( c ), mixed forest ( d ), grassland ( e ), shrubland ( f ), and savanna ( g ) across continents estimated by GPP, kNDVI, and EVI, respectively. Boxplots show the statistics – minima: first interquartile subtract 1.5 times the interquartile range, maxima: third interquartile add 1.5 times the interquartile range, horizontal lines: median; dots: mean; boxes: interquartile range; whiskers: 1.5 times the interquartile range. The interquartile range denote the difference between the third and first interquartile. The numbers in parentheses represent the sample sizes. The recovery time was compared between continent pairs using the Student’s t-test, and the connected line in the subplot denotes when there were no significant differences (P > 0.01) for the same vegetation type in two continents, while unconnected two continents denote that there were significant differences (P < 0.01) in the recovery time among the same vegetation type in the two continents. We excluded continents when there were less than 500 burned and recovered pixels available for a specific vegetation type. ns denotes not significant.

Extended Data Fig. 5 Spatial pattern of average recovery time (estimated by kNDVI), and statistical profiles of recovery time in vegetation types, hydrothermal space.

a , Spatial distribution of mean recovery time. The inset depicts the area fraction of different recovery times. b , Recovery time change along latitude gradients. Shaded areas are the 95% confidence intervals for the average recovery time. c , Violin plot illustrating recovery times across various vegetation types. Boxplots show the statistics – minima: first interquartile subtract 1.5 times the interquartile range, maxima: third interquartile add 1.5 times the interquartile range, horizontal lines: median; dots: mean; boxes: interquartile range; whiskers: 1.5 times the interquartile range. The interquartile range denote the difference between the third and first interquartile. The numbers in parentheses represent the sample sizes. d , Variation in average recovery time as a function of annual mean air temperature and precipitation. Map a created in ArcGIS (v. 10.8) using base map derived from the Global Self-consistent, Hierarchical, High-resolution Geography Database ( https://www.ngdc.noaa.gov/mgg/shorelines/data/gshhg/latest/ ).

Extended Data Fig. 6 Spatial pattern of average recovery time (estimated by EVI), and statistical profiles of recovery time in vegetation types, hydrothermal space.

Extended data fig. 7 recovery time comparison between intact and modified landscapes using the student’s t-test..

a–c , The recovery time estimated by GPP (a), kNDVI ( b ), and EVI ( c ), respectively. Boxplots show the statistics – minima: first interquartile subtract 1.5 times the interquartile range, maxima: third interquartile add 1.5 times the interquartile range, horizontal lines: median; dots: mean; boxes: interquartile range; whiskers: 1.5 times the interquartile range. The interquartile range denote the difference between the third and first interquartile. The numbers in parentheses represent the sample sizes. The modified landscapes were generated near the intact landscapes within a 50 km buffer.

Extended Data Fig. 8 Recovery trajectory for different vegetation types.

a , Spatial distribution of vegetation types. b–m , Examples of recovery trajectory for a pixel and fire event in ( b ) grassland in North America, ( c ) deciduous broadleaf forest in Europe, ( d ) grassland in central Asia, ( e ) evergreen needleleaf forest in Russia, ( f ) savanna in Asia, ( g ) deciduous broadleaf forest in Asia, ( h ) deciduous broadleaf forest in southern North America, ( i ) shrubland in South America, ( j ) deciduous broadleaf forest in Africa, ( k ) grassland in Africa, ( l ) evergreen broadleaf forest in Southeast Asia, ( m ) grassland in Australia. The post-fire vegetation productivity recovery in ( d ), ( g ), ( i ), and ( m ) was dominated by Rec SM , in ( b ), ( e ), ( h ), ( j ), ( k ), and ( i ) by Rec T , in ( c ) and ( f ) by Rec VPD . The magenta bar indicates fire disturbance, whereas the shaded vertical bars represent the intensity of the fire disturbance. The larger the burning area, the darker the magenta shading. For each typical region, the dominant factor on recovery time was derived from the results of Fig. 3 . The fire response lag denotes the duration of months between when the fire occurred and when the vegetation productivity reached the maximum loss. Map a created in ArcGIS (v. 10.8). Map a created in ArcGIS (v. 10.8) using base maps data derived from the ESA Climate Change Initiative ( https://www.esa-landcover-cci.org/?q=node/164 ) and the Global Self-consistent, Hierarchical, High-resolution Geography Database ( https://www.ngdc.noaa.gov/mgg/shorelines/data/gshhg/latest/ ).

Extended Data Fig. 9 Spatial distribution of the land cover changes before and after fires estimated using the burned and recovered pixels.

a–c , The recovery time estimated by GPP ( a ), kNDVI ( b ), and EVI ( c ), respectively. The inserted histograms show the proportions of each type of land cover change. Map a created in ArcGIS (v. 10.8) using base map derived from the Global Self-consistent, Hierarchical, High-resolution Geography Database ( https://www.ngdc.noaa.gov/mgg/shorelines/data/gshhg/latest/ ).

Supplementary information

Supplementary information.

Supplementary Figs. 1–17 and Tables 1 and 2.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Cite this article.

Xu, H., Chen, H.W., Chen, D. et al. Global patterns and drivers of post-fire vegetation productivity recovery. Nat. Geosci. (2024). https://doi.org/10.1038/s41561-024-01520-3

Download citation

Received : 08 January 2024

Accepted : 24 July 2024

Published : 23 August 2024

DOI : https://doi.org/10.1038/s41561-024-01520-3

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

research paper on fire ecology

Frontiers in Fire Ecology

New Content Item

Edited by: Karin L. Riley, Leda Kobziar, Andrew T. Hudak

Association for Fire Ecology celebrated its 20th anniversary in 2017 at the 7th International Fire Ecology and Management Congress . More than 350 research presentations highlighted advances in fire ecology and fire management science, policy, and practice. A select group of presentations was identified by the Editors of this collection as being especially integral to the advancement of our discipline; presenters were invited to submit manuscripts for the Association’s Frontiers in Fire Ecology collection.

Not only does this compilation of articles reflects an ever-increasing understanding of the role of fire in our world today, it also celebrates the novel contributions of young scientists in charting the future of fire ecology and management. Together, these manuscripts reflect advances in technology, theory, ecology, modeling, workforce social dynamics, and the application of each of these to furthering the integration of fire as a driving force into the human-environment relationship.

Frontiers in Fire Ecology represents current advances and directions for the future of fire ecology and related research.

Nitrogen allocation in PM 2.5 smoke-exposed plants: implications for ecosystem nitrogen cycling and stress response

With the increase in forest fire emissions, an increasing amount of nitrogen is released from combustibles and taken up by plant leaves in the form of PM 2.5 smoke deposition. Concurrently, the stress from PM 2.5 a...

  • View Full Text

research paper on fire ecology

Smoke in the Great Plains, USA: an increasing phenomenon with potential policy and health implications

In recent decades, as wildland fire occurrence has increased in the United States, concern about the emissions produced by wildland fires has increased as well. This growing concern is evidenced by an increase...

Prescribed fire science: the case for a refined research agenda

The realm of wildland fire science encompasses both wild and prescribed fires. Most of the research in the broader field has focused on wildfires, however, despite the prevalence of prescribed fires and demons...

Empowering strategic decision-making for wildfire management: avoiding the fear trap and creating a resilient landscape

In recent years, fire services in Mediterranean Europe have been overwhelmed by extreme wildfire behavior. As a consequence, fire management has moved to defensive strategies with a focus only on the known ris...

Long-term vegetation response following post-fire straw mulching

Straw mulching is one of the most common treatments applied immediately post fire to reduce soil erosion potential and mitigate post-fire effects on water quality, downstream property, and infrastructure, but ...

Characterizing persistent unburned islands within the Inland Northwest USA

In the Inland Pacific Northwest of the United States, fire is a dominant driver of ecological change. Within wildfire perimeters, fire effects often vary considerably and typically include remnant patches of u...

Tree regeneration following wildfires in the western US: a review

Wildfires, like many disturbances, can be catalysts for ecosystem change. Given projected climate change, tree regeneration declines and ecosystem shifts following severe wildfires are predicted. We reviewed s...

Got shrubs? Precipitation mediates long-term shrub and introduced grass dynamics in chaparral communities after fire

Short-term post-fire field studies have shown that native shrub cover in chaparral ecosystems negatively affects introduced cover, which is influenced by burn severity, elevation, aspect, and climate. Using th...

Examining post-fire vegetation recovery with Landsat time series analysis in three western North American forest types

Few studies have examined post-fire vegetation recovery in temperate forest ecosystems with Landsat time series analysis. We analyzed time series of Normalized Burn Ratio (NBR) derived from LandTrendr spectral...

Fuels, vegetation, and prescribed fire dynamics influence ash production and characteristics in a diverse landscape under active pine barrens restoration

An important consequence of wildland fire is the production of ash, defined as a continuum of mineral to charred organic residues formed by the burning of wildland fuels. Ash may impact soil health depending o...

The Correction to this article has been published in Fire Ecology 2020 16 :7

Restoring historical fire regimes increases activity of endangered bats

Fire suppression has altered ecological communities globally. Prescribed fire regimes strive to restore function to these fire-dependent ecosystems by mimicking natural fire regimes. Although fire frequency is...

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

fire-logo

Article Menu

research paper on fire ecology

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Current status of research on wildland fire impacts on soil environment and soil organisms and hotspots visualization analysis.

research paper on fire ecology

Graphical Abstract

1. Introduction

2. data sources and methods, 3.1. current status of research on the impact of wildland fire on the soil environment, 3.1.1. state of the art, 3.1.2. annual publication volume, countries, institutions, and authors, 3.1.3. keyword co-occurrence and clustering, 3.2. current research status of the impact of wildland fire on soil organisms, 3.2.1. state of the art, 3.2.2. annual publication volume, countries, institutions, and authors, 3.2.3. keyword co-occurrence and clustering for wildland fire effects on soil, microorganisms, 3.3. current research status of the impact of wildland fire on soil fauna, 3.3.1. state of the art, 3.3.2. annual publication volume, countries, institutions, and authors, 3.3.3. keyword co-occurrence and clustering for the impact of wildland fire on soil fauna, 4. conclusions, 5. existing problems and perspectives, author contributions, data availability statement, acknowledgments, conflicts of interest.

  • Hart, S.; Deluca, T.H.; Newman, G.S.; Mackenzie, M.D.; Boyle, S.I. Post-fire vegetative dynamics as drivers of microbial community structure and function in forest soils. For. Ecol. Manag. 2005 , 220 , 166–184. [ Google Scholar ]
  • Alcañiz, M.; Outeiro, L.; Francos, M.; Ubeda, X. Effects of prescribed fires on soil properties: A review. Sci. Total Environ. 2018 , 613 , 944–957. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Keeley, J.E. Fire intensity, fire severity and burn severity: A brief review and suggested usage. Int. J. Wildland Fire 2009 , 18 , 116–126. [ Google Scholar ] [ CrossRef ]
  • Xue, L.; Li, Q.; Chen, H. Effects of a Wildfire on Selected Physical, Chemical and Biochemical Soil Properties in a Pinus massoniana Forest in South China. Forests 2014 , 5 , 2947–2966. [ Google Scholar ] [ CrossRef ]
  • Cheng, Z.; Wu, S.; Du, J.; Pan, H.; Lu, X.; Liu, Y.; Yang, L. Variations in the Diversity and Biomass of Soil Bacteria and Fungi under Different Fire Disturbances in the Taiga Forests of Northeastern China. Forests 2023 , 14 , 2063. [ Google Scholar ] [ CrossRef ]
  • Perez-Valera, E.; Goberna, M.; Faust, K.; Raes, J.; Garcia, C.; Verdu, M. Fire modifies the phylogenetic structure of soil bacterial co-occurrence networks. Environ. Microbiol. 2017 , 19 , 317–327. [ Google Scholar ] [ PubMed ]
  • Espinosa, J.; Dejene, T.; Guijarro, M.; Cerdá, X.; Madrigal, J.; Martín-Pinto, P. Fungal diversity and community composition responses to the reintroduction of fire in a non-managed Mediterranean shrubland ecosystem. For. Ecosyst. 2023 , 10 , 100110. [ Google Scholar ] [ CrossRef ]
  • Qiang, W.; He, L.; Zhang, Y.; Liu, B.; Pang, X. Aboveground vegetation and soil physicochemical properties jointly drive the shift of soil microbial community during subalpine secondary succession in southwest China. Catena 2021 , 202 , 105251. [ Google Scholar ] [ CrossRef ]
  • Brown, G.; Edwards, C.; Brussaard, L. How earthworms affect plant growth: Burrowing into the mechanisms. Earthworm Ecol. 2004 , 2 , 13–49. [ Google Scholar ] [ CrossRef ]
  • Lussenhop, J. Mechanisms of microarthropod-microbial interactions in soil. Adv. Ecol. Res. 1992 , 23 , 1–33. [ Google Scholar ] [ CrossRef ]
  • Yang, S.J.; Zhang, M.T.; Wei, H.; Zhang, B.; Peng, J.; Shang, P.F.; Sun, S.K. Research prospects for kidney xenotransplantation: A bibliometric analysis. Ren. Fail. 2024 , 46 , 2301681. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Nurhidayah, N.; Sudarma, M.; Atmini, S. Audit opinion research: Overview and research agenda. Cogent Bus. Manag. 2024 , 11 , 2301134. [ Google Scholar ] [ CrossRef ]
  • Zhang, G.F.; Qin, Y.M.; Liu, S.B. Bibliometric analysis of research trends and topic areas in traditional Chinese medicine therapy for lymphoma. Pharm. Biol. 2023 , 62 , 12–31. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hu, X.Q.; Ji, Z.X.; Gu, S.H.; Ma, Z.R.; Yan, Z.S.; Liang, Y.; Chang, H.Q.; Liang, H. Mapping the research on desulfurization wastewater: Insights from a bibliometric review (1991–2021). Chemosphere 2023 , 314 , 137678. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Do Carmo, G.; Felizardo, L.F.; Alcantara, V.D.; da Silva, C.A.; do Prado, J.W. The impact of Jurgen Habermas’s scientific production: A scientometric review. Scientometrics 2023 , 128 , 1853–1875. [ Google Scholar ] [ CrossRef ]
  • Tsoulfas, G.T.; Trivellas, P.; Reklitis, P.; Anastasopoulou, A. A Bibliometric Analysis of Short Supply Chains in the Agri-Food Sector. Sustainability 2023 , 15 , 1089. [ Google Scholar ] [ CrossRef ]
  • Ding, Q.; Song, X.L.; Pu, Y.P. Removal of microcystins from water and primary treatment technologies A comprehensive understanding based on bibliometric and content analysis, 1991–2020. J. Environ. Manag. 2022 , 305 , 114349. [ Google Scholar ] [ CrossRef ]
  • Wang, S.; Xia, D.M.; Zhang, Z.T.; Zhang, J.L.; Meng, W.H.; Zhang, Y.P.; Xu, S.G. Mapping Trends and Hotspots Regarding the Use of Ultrasound in Emergency Medicine: A Bibliometric Analysis of Global Research. Front. Public Health 2021 , 9 , 764642. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Huang, Y.-J.; Cheng, S.; Yang, F.-Q.; Chen, C. Analysis and Visualization of Research on Resilient Cities and Communities Based on VOSviewer. Int. J. Environ. Res. Public Health 2022 , 19 , 7068. [ Google Scholar ] [ CrossRef ]
  • Hatten, J.A.; Zabowski, D.; Ogden, A.; Thies, W. Soil organic matter in a ponderosa pine forest with varying seasons and intervals of prescribed burn. For. Ecol. Manag. 2008 , 255 , 2555–2565. [ Google Scholar ] [ CrossRef ]
  • Dai, Q.; Zhu, J.; Lv, G.; Kalin, L.; Yao, Y.; Zhang, J.; Han, D. Radar remote sensing reveals potential underestimation of rainfall erosivity at the global scale. Sci. Adv. 2023 , 9 , eadg5551. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jia, B.R. Litter decomposition and its underlying mechanisms. Chin. J. Plant Ecol. 2019 , 43 , 648–657. [ Google Scholar ] [ CrossRef ]
  • Wang, S.S.; Long, J.; Ding, H. Leaf water absorption and canopy rainfall interception of twenty-one plant species in Beijing. J. Beijing For. Univ. 2020 , 42 , 100–110. [ Google Scholar ]
  • Liu, F.L.; Chen, X.W.; Zeng, S.P.; Peng, Z.Z. Progress of the effects of fire disturbance on forest soil water repellency. Acta Ecol. Sin. 2019 , 39 , 1846–1852. [ Google Scholar ] [ CrossRef ]
  • Carter, M.C.; Foster, C.D. Prescribed burning and productivity in southern pine forests: A review. For. Ecol. Manag. 2004 , 191 , 93–109. [ Google Scholar ]
  • Cheng, C.H.; Chen, Y.S.; Huang, Y.H.; Chiou, C.R.; Lin, C.C.; Menyailo, O.V. Effects of repeated fires on ecosystem C and N stocks along a fire induced forest/grassland gradient. J. Geophys. Res. Biogeosciences 2013 , 118 , 215–225. [ Google Scholar ] [ CrossRef ]
  • Hamman, S.T.; Burke, I.C.; Knapp, E.E. Soil Nutrients and Microbial Activity after Early and Late Season Prescribed Burns in a Sierra Nevada Mixed Conifer Forest. For. Ecol. Manag. 2008 , 256 , 367–374. [ Google Scholar ]
  • Zeng, S.P.; Liu, F.L.; Zhao, M.F.; Wang, G.J.; Chen, X.W. Effects of fire disturbance intensities on soil physiochemical properties of pour subtropical forest types. Acta Ecol. Sin. 2020 , 40 , 233–246. [ Google Scholar ] [ CrossRef ]
  • Butler, O.M.; Elser, J.J.; Lewis, T.; Mackey, B.; Chen, C.R. The phosphorus-rich signature of fire in the soil-plant system: A global meta-analysis. Ecol. Lett. 2018 , 21 , 335–344. [ Google Scholar ] [ CrossRef ]
  • Wang, F.M.; Li, J.; Zou, B.; Xu, X.; Li, Z.A. Effect of Prescribed Fire on Soil Properties and N Transformation in Two Vegetation Types in South China. Environ. Manag. 2013 , 51 , 1164–1173. [ Google Scholar ] [ CrossRef ]
  • Hu, H.Q.; Luo, S.S.; Luo, B.Z.; Wei, S.J.; Wu, Z.P.; Wang, Z.S.; Li, X.C.; Zhou, Y.F. Effects of forest fire disturbance on soil organic carbon in forest ecosystems: A review. Acta Ecol. Sin. 2020 , 40 , 839–1850. [ Google Scholar ] [ CrossRef ]
  • Urbanski, S. Wildland fire emissions, carbon, and climate: Emission factors. For. Ecol. Manag. 2014 , 317 , 51–60. [ Google Scholar ]
  • Hume, A.; Chen, H.Y.H.; Taylor, A.R.; Kayahara, G.J.; Man, R.Z. Soil C:N:P dynamics during secondary succession following fire in the boreal forest of central Canada. Forest Ecol. Manag. 2016 , 369 , 1–9. [ Google Scholar ]
  • Xu, P.B.; Qu, M.; Xue, L. Effects of forest fire on forest soils. Chin. J. Ecol. 2013 , 32 , 1596–1606. [ Google Scholar ] [ CrossRef ]
  • Sheikh, A.; Siddique, N.; Qutab, S.; Khan, M.A.; Mahmood, K. An investigation of emerging COVID-19 research trends and future implications for LIS field: A bibliometric mapping and visualization. J. Librariansh. Inf. Sci. 2023 , 55 , 3–17. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hernández-González, O.; Fresno-Rodríguez, A.; Spencer-Contreras, R.E.; Tárraga-Mínguez, R.; González-Fernández, D.; Sepúlveda-Opazo, F. Research Mapping of Trauma Experiences in Autism Spectrum Disorders: A Bibliometric Analysis. Healthcare 2023 , 11 , 1267. [ Google Scholar ] [ CrossRef ]
  • Boshoff, N.; Ngwenya, S. Agricultural Research in Zimbabwe: An Author-level Bibliometric Analysis of Publication Outlets and Research Collaboration. Sci. Technol. Soc. 2022 , 27 , 404–428. [ Google Scholar ] [ CrossRef ]
  • Ma, W.; Opp, C.; Yang, D. Past, Present, and Future of Virtual Water and Water Footprint. Water 2020 , 12 , 3068. [ Google Scholar ] [ CrossRef ]
  • Li, J.; Mao, Y.; Ouyang, J.; Zheng, S. A Review of Urban Microclimate Research Based on CiteSpace and VOSviewer Analysis. Int. J. Environ. Res. Public Health 2022 , 19 , 4741. [ Google Scholar ] [ CrossRef ]
  • Cui, M.Q.; Wu, C.; Jiang, X.X.; Liu, Z.Y.; Xue, S.G. Bibliometric analysis of research on soil arsenic during 2005–2016. J. Cent. South Univ. 2019 , 26 , 479–488. [ Google Scholar ] [ CrossRef ]
  • Maléchaux, A.; Le Dréau, Y.; Artaud, J.; Dupuy, N. Exploring the Scientific Interest for Olive Oil Origin: A Bibliometric Study from 1991 to 2018. Foods 2020 , 9 , 556. [ Google Scholar ] [ CrossRef ]
  • Peset Mancebo, M.F.; Garzón Farinós, M.F.; Gonzalez, L.; García-Massó, X.; Ferrer Sapena, A.; Toca-Herrera, J.L.; Sánchez Pérez, E.A. Survival analysis of author keywords: An application to the library and information sciences area. J. Assoc. Inf. Sci. Technol. 2020 , 71 , 462–473. [ Google Scholar ] [ CrossRef ]
  • Li, F.; Shi, Z.; Zhao, B.; Bono, G.J.; Sun, L.; Hu, T. The Effects of Fire Disturbance on Litter Decomposition and C:N:P Stoichiometry in a Larix gmelinii Forest Ecosystem of Boreal China. Forests 2022 , 13 , 1029. [ Google Scholar ] [ CrossRef ]
  • Liu, X.; Pan, C. Effects of recovery time after fire and fire severity on stand structure and soil of larch forest in the Kanas National Nature Reserve, Northwest China. J. Arid Land 2019 , 11 , 811–823. [ Google Scholar ] [ CrossRef ]
  • Lee, C.W.; Seo, J.I.; Youn, H.J.; Kim, K. Effectiveness of rehabilitation treatments on a slowly revegetating hillslope in a recently burned coastal forest, Republic of Korea. Landsc. Ecol. Eng. 2014 , 10 , 249–260. [ Google Scholar ] [ CrossRef ]
  • Aparício, B.A.; Santos, J.A.; Freitas, T.R.; Sá, A.C.L.; Pereira, J.M.C.; Fernandes, P.M. Unravelling the effect of climate change on fire danger and fire behaviour in the Transboundary Biosphere Reserve of Meseta Iberica (Portugal-Spain). Clim. Change 2022 , 173 , 5. [ Google Scholar ] [ CrossRef ]
  • Lucas-Borja, M.E.; Plaza-Alvarez, P.A.; Gonzalez-Romero, J.; Sagra, J.; Alfaro-Sánchez, R.; Zema, D.A.; Moya, D.; de las Heras, J. Short-term effects of prescribed burning in Mediterranean pine plantations on surface runoff, soil erosion and water quality of runoff. Sci. Total Environ. 2019 , 674 , 615–622. [ Google Scholar ] [ CrossRef ]
  • Liu, W.C.; Zhang, Z.; Li, J.X.; Wen, Y.H.; Liu, F.H.; Zhang, W.; Liu, H.Y.; Ren, C.J.; Han, X.H. Effects of fire on the soil microbial metabolic quotient: A global meta-analysis. Catena 2023 , 224 , 106957. [ Google Scholar ] [ CrossRef ]
  • Abreu, C.I.; Friedman, J.; Woltz, V.L.A.; Gore, J. Mortality causes universal changes in microbial community composition. Nat. Commun. 2019 , 10 , 2120. [ Google Scholar ] [ CrossRef ]
  • Wu, X.M.; Ma, L.; Tao, Y.Z.; Di, X.Y.; Zhang, X.Y. Instant Impact of Fire on Birch Soil Microbial in Tahe Larch Forest. J. Northeast For. Univ. 2015 , 43 , 84–87. [ Google Scholar ] [ CrossRef ]
  • Fultz, L.M.; Moore-Kucera, J.; Dathe, J.; Davinic, M.; Perry, G.; Wester, D.; Schwilk, D.W.; Rideout-Hanzak, S. Forest wildfire and grassland prescribed fire effects on soil biogeochemical processes and microbial communities: Two case studies in the semi-arid Southwest. Appl. Soil Ecol. 2016 , 99 , 118–128. [ Google Scholar ] [ CrossRef ]
  • Vild, O.; Kalwij, J.M.; Hédl, R. Effects of simulated historical tree litter raking on the understorey vegetation in a central European forest. Appl. Veg. Sci. 2015 , 18 , 569–578. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Chen, L.Y.; Liu, L.; Qin, S.Q.; Yang, G.B.; Fang, K.; Zhu, B.; Kuzyakov, Y.; Chen, P.D.; Xu, Y.P.; Yang, Y.H. Regulation of priming effect by soil organic matter stability over a broad geographic scale. Nat. Commun. 2019 , 10 , 5112. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Fierer, N.; Bradford, M.A.; Jackson, R.B. Toward an ecological classification of soil bacteria. Ecology 2007 , 88 , 1343–1344. [ Google Scholar ] [ CrossRef ]
  • Mataix-Solera, J.; Guerrero, C.; García-Orenes, F.; Bárcenas-Moreno, G.; Torres, M.P. Fire effects on soils and restoration strategies. For. Fire Eff. Soil Microbiol. 2009 . [ Google Scholar ] [ CrossRef ]
  • Bárcenas-Moreno, G.; Bååth, E. Bacterial and fungal growth in soil heated at different temperatures to simulate a range of fire intensities. Soil Biol. Biochem. 2009 , 41 , 2517–2526. [ Google Scholar ] [ CrossRef ]
  • Cheng, Z.; Wu, S.; Pan, H.; Lu, X.; Liu, Y.; Yang, L. Cortinarius and Tomentella Fungi Become Dominant Taxa in Taiga Soil after Fire Disturbance. J. Fungi. 2023 , 9 , 1113. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cheng, Z.; Wu, S.; Du, J.; Liu, Y.; Sui, X.; Yang, L. Reduced Arbuscular Mycorrhizal Fungi (AMF) Diversity in Light and Moderate Fire Sites in Taiga Forests, Northeast China. Microorganisms 2023 , 11 , 1836. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Baldrian, P.; López-Mondéjar, R.; Kohout, P. Forest microbiome and global change. Nat. Rev. Microbiol. 2023 , 21 , 487–501. [ Google Scholar ] [ CrossRef ]
  • Nelson, A.R.; Narrowe, A.B.; Rhoades, C.C.; Fegel, T.S.; Daly, R.A.; Roth, H.K.; Chu, R.K.; Amundson, K.K.; Young, R.B.; Steindorff, A.S.; et al. Wildfire-dependent changes in soil microbiome diversity and function. Nat. Microbiol. 2022 , 7 , 1419–1430. [ Google Scholar ] [ CrossRef ]
  • Hu, H.Q. Forest Fire Ecology and Management ; China Forestry Publishing House: Beijing, China, 2005. [ Google Scholar ]
  • Wiens, J.A. Spatial Scaling in Ecology. Funct. Ecol. 1989 , 3 , 385. [ Google Scholar ] [ CrossRef ]
  • Hobbs, R.J. Future landscapes and the future of landscape ecology. Landsc. Urban Plan. 1997 , 37 , 1–9. [ Google Scholar ] [ CrossRef ]
  • Zhang, S.H.; Zhang, X.P. Guilds of Soil Macro-animals in Burned Forest Areas in Daxing’an Mountains. J. Northeast For. Univ. 2011 , 39 , 70–72. [ Google Scholar ] [ CrossRef ]
  • Chen, W.K. The Diversity of Nematodes in Aciculiailvae Burned Area of Alpine Region in Batang County ; Chengdu University of Technology: Chengdu, China, 2016. [ Google Scholar ]
  • Shao, Y.H.; Wang, Z.Y.; Liu, T.; Kardol, P.; Ma, C.G.; Hu, Y.H.; Cui, Y.; Zhao, C.C.; Zhang, W.X.; Guo, D.L.; et al. Drivers of nematode diversity inforest soils across climatic zones. Proc. R. Soc. B 2023 , 290 , 20230107. [ Google Scholar ] [ CrossRef ]
  • Gongalsky, K.B.; Persson, T.; Pokarzhevskii, A.D. Effects of soil temperature and moisture on the FeedIng Activity of soil animals as determined by the bait-lamina test. Appl. Soil Ecol. 2008 , 39 , 84–90. [ Google Scholar ] [ CrossRef ]
  • Gongalsky, K.B.; Malmström, A.; Zaitsev, A.S.; Shakhab, S.V.; Bengtsson, J.; Persson, T. Do burned areas recover from inside? An experiment with soil fauna in a heterogeneous landscape. Appl. Soil Ecol. 2012 , 59 , 73–86. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

AffiliationRecord CountAffiliationRecord Count
United States Department of Agriculture157Consejo Superior de Investigaciones Cientificas87
United States Forest Service116Chinese Academy of Sciences49
United States Department of Interior101University of Idaho46
United States Geological Survey96Russian Academy of Sciences42
University of California88Swansea University41
No.KeywordOccurrenceNo.KeywordOccurrence
1Wildfire6956Vegetation184
2Fire5027Climate Change173
3Organic matter2488Carbon172
4Soil2149Erosion162
5Nitrogen18510Forest159
AffiliationRecord CountAffiliationRecord Count
University of California 45Oregon State University18
United States Department of Agriculture 40United States Department of Energy18
United States Forest Service34Chinese academy of Sciences16
Consejo Superior de Investigaciones Cientificas27Swedish University of Agricultural Sciences15
Universidad de Valladolid19Czech Academy of Sciences14
No.KeywordOccurrenceNo.KeywordOccurrence
1Wildfire2146Soil69
2Fire1827Nitrogen69
3Diversity1308Dynamics65
4Organic matter829Forest62
5Carbon8110Biomass60
AffiliationRecord CountAffiliationRecord Count
United States Department of Agriculture12Swedish University of Agricultural Sciences7
Russian Academy of Sciences12United States Department of Interior6
Severtsov Institute of Ecology and Evolution8United States Geological Survey6
Saratov Scientific Center of the Russian Academy of Sciences8Australian National University6
United States Forest Service8Justus Liebig University Giessen5
No.KeywordOccurrencesNo.KeywordOccurrences
1Wildfire516Disturbance16
2Fire477Diversity15
3Vegetation218Management15
4Biodiversity209Carbon13
5Response1710Climate Change13
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Cheng, Z.; Wu, S.; Wei, D.; Pan, H.; Fu, X.; Lu, X.; Yang, L. Current Status of Research on Wildland Fire Impacts on Soil Environment and Soil Organisms and Hotspots Visualization Analysis. Fire 2024 , 7 , 163. https://doi.org/10.3390/fire7050163

Cheng Z, Wu S, Wei D, Pan H, Fu X, Lu X, Yang L. Current Status of Research on Wildland Fire Impacts on Soil Environment and Soil Organisms and Hotspots Visualization Analysis. Fire . 2024; 7(5):163. https://doi.org/10.3390/fire7050163

Cheng, Zhichao, Song Wu, Dan Wei, Hong Pan, Xiaoyu Fu, Xinming Lu, and Libin Yang. 2024. "Current Status of Research on Wildland Fire Impacts on Soil Environment and Soil Organisms and Hotspots Visualization Analysis" Fire 7, no. 5: 163. https://doi.org/10.3390/fire7050163

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

  • Search Menu
  • Sign in through your institution
  • Advance articles
  • Editor's Choice
  • Special Collections
  • Author Guidelines
  • Submission Site
  • Open Access
  • Reasons to submit
  • About BioScience
  • Journals Career Network
  • Editorial Board
  • Advertising and Corporate Services
  • Self-Archiving Policy
  • Potentially Offensive Content
  • Terms and Conditions
  • Journals on Oxford Academic
  • Books on Oxford Academic

Issue Cover

Article Contents

Early evolutionary fire ecology: fauna, early evolutionary fire ecology: flora, fire enlightens evolutionary biology, future directions in evolutionary fire ecology, acknowledgments, author contributions, references cited.

  • < Previous

Evolutionary fire ecology: An historical account and future directions

ORCID logo

  • Article contents
  • Figures & tables
  • Supplementary Data

Juli G Pausas, Jon E Keeley, Evolutionary fire ecology: An historical account and future directions, BioScience , Volume 73, Issue 8, August 2023, Pages 602–608, https://doi.org/10.1093/biosci/biad059

  • Permissions Icon Permissions

The idea that fire acts as an evolutionary force contributing to shaping species traits started a century ago, but had not been widely recognized until very recently. Among the first to realize this force were E dward B . Poulton, R. D ale Guthrie, and E dwin V . Komarek in animals and W illis L . Jepson, W alter W . Hough, T om M . Harris, P hilip V . Wells, and R obert W . Mutch in plants. They were all ahead of their time in their evolutionary thinking. Since then, evolutionary fire ecology has percolated very slowly into the mainstream ecology and evolutionary biology; in fact, this topic is still seldom mentioned in textbooks of ecology or evolution. Currently, there is plenty of evidence suggesting that we cannot understand the biodiversity of our planet without considering the key evolutionary role of fire. But there is still research to be done in order to fully understand fire's contribution to species evolution and to predicting species responses to rapid global changes.

Nothing in fire science makes sense except in the light of evolution. (paraphrased from Dobzhansky 1973 ).

Our understanding of evolution by natural selection is largely a legacy of Charles Darwin. As a great naturalist, he observed a myriad of evidence of natural selection when traveling around the world. However, there is no evidence that he ever thought of fire as a natural phenomenon acting as a selective pressure, despite exploring fire-prone ecosystems in South America and Australia (Nicholas and Nicholas 2008 ). Here are some notes in the HMS Beagle diary (Keynes 2001 ) that show his contact with fire-prone ecosystems:

“ We have seen during the day the smoke from several large fires within the country: It is not easy to guess how they arise. It is too far North for the Indians and the country is uninhabited by the Spaniards ” (Keynes 2001 : 95; 24 August 1832, Buenos Aires province, Argentina).

“ After having crossed the monotonous Savannahs of grass, the gardens and Orchards around the town are very pleasing ” (Keynes 2001 : 316; 28 March 1835, Mendoza, Argentina).

“ In the whole country I scarcely saw a place without the marks of a fire; whether these had been more or less recent—whether the stumps were more or less black, was the greatest change which varied the uniformity so wearisome of the traveler's eye ” (Keynes 2001 : 402; 19 January 1836, New South Wales, Australia).

“ We passed through large tracts of country in flames; volumes of smoke sweeping across the road and that I scarcely saw a place, without the marks of fires ” (Keynes 2001 : 405; 23 January 1836, New South Wales, Australia).

Alexander von Humboldt, another great naturalist, also visited fire-prone ecosystems of South America (savannas) a few years earlier than Darwin and considered savannas as deforested environments (Pausas and Bond 2019 ). An evolutionary role for fire never occurred to these early explorers and naturalists because they carried a cultural bias of looking at the world through the view of Central European foresters, who considered fire as a human factor (Pausas and Bond 2019 ).

We have learned a lot since the early naturalists, and currently, there is overwhelming evidence that fires are very old in the geological history of the planet (Scott 2018 ) and that many plants and animals have adapted to the particular fire regimes they have been subject to through their history (Keeley et al. 2011 , 2012 , He et al. 2012 , Simon and Pennington 2012 , Charles-Dominique et al. 2015 , Pausas 2015a , b , Pausas and Parr 2018 , Lamont et al. 2019 , 2020 , Keeley and Pausas 2022 ). In the present article, we aim to make a tribute to early scientists (primarily earlier than 1970) who were ahead of their time by understanding the evolutionary force of fire, in an era when most were unconvinced or anathema to recognizing fire as a selective factor in trait evolution. These early scientists are a key part of the history of biology. Below, we ask who were those earliest researchers suggesting that fire could act as an evolutionary force generating adaptations. By doing so, we provide the first historical approach on the evolutionary adaptations to fire. Then we briefly mention why this is so important, and we end by outlining possible future directions.

In searching for old references, we focused on those published in English, because it is the language with the most information on the topic (and in science in general), although we have reviewed some papers in Romance languages (e.g., Pausas 1997 ). However, the possibility exists that we have missed some old references on evolutionary fire ecology from non-English researchers; if so, the authors would appreciate feedback on such omissions. And as you will see below, all early researchers with an evolutionary perspective of fire we are mentioning are male. The invisibility of women scientists is a long-standing problem; we would also appreciate very much feedback on any female scientists we may have omitted.

Edward B. Poulton (1856–1943) was an evolutionary biologist who was a strong supporter of Darwin; he was probably among the first to think that fire could generate an evolutionary pressure. And this idea came to him when investigating fire beetles (Buprestids, Coleoptera)—that is, wood-boring beetles that evolved infrared sensors to locate recently burned areas (as mating sites and an ideal environment for reproductive success).

“ Poulton said that the instinct of the beetle, like the wonderful fire-resisting powers of many Australian trees, had probably been developed in ancient times as a response to bush fires ” (Poulton 1915 : iv).

It is interesting that he compared these observations of fire beetles with Australian trees, so he perhaps also thought that fire-resisting traits in plants were fire adaptations. In 1926, after his visit to Africa, he suggested another fire adaptation in animals—specifically, the adaptation of some insects to the postfire environment (fire mimicry):

“ Another difference often observable follows from the fact that a grassfire sweeps rapidly through the dry growth and leaves the stronger stalks scorched and charred but standing. Many species are adapted to this environment not by developing a melanic form, but one in which the black and darkened straw colour are combined ” (Poulton 1926 ).

Increased postfire melanism was also proposed as an adaptive response to postfire environments (camouflage) in ground squirrels by R. Dale Guthrie (born 1936; Guthrie 1967 ). In a review of the role of fire on animal behavior, Edwin V. Komarek Jr. (1905–1995) concluded that

“ it is evident that many animals are adapted to a fire environment and that natural selection has been a major factor in such adaptation ” (Komarek 1969 ).

Despite those early observations (and others a few years later, e.g., Lillywhite et al. 1977 ), little has been done to demonstrate the role of fire as generating adaptations in the animal kingdom (Pausas and Parr 2018 ). Certainly, there are many examples of animals that are more abundant after a fire or in fire-prone ecosystems, suggesting that they are likely adapted to fire environments (e.g., Hutto 2008 ; for a review, see Pausas and Parr 2018 ), but few studies on animals have used an evolutionary approach in relation to fire. The mobility of animals suggests many of the adaptations to be behavioral rather than morphological (as in plants) and, therefore, more difficult to tie to fire (Pausas and Parr 2018 ). Apparently, early biologists were not aware of those potential behavioral adaptations; only recently, there has been an increase in behavioral studies testing the sensitivity of animals to fire cues (e.g., smoke, fire sound; Stromberg 1997 , Stawski et al. 2015 , Nowack et al. 2018 , Álvarez-Ruiz et al. 2021 , Nimmo et al. 2021 ). Among recent papers that have been focused on fire melanism, a key example is the study of natural selection to postfire conditions in grasshoppers (Forsman et al. 2011 ); apparently, no such studies have been done in vertebrates.

The oldest published mention of a fire adaptation in plants that we are aware of is by Willis L. Jepson (1867–1946) in discussing the role of fire in the life cycle of the California chaparral shrub Arctostaphylos sensitiva (Jepson 1922 ). This is a postfire obligate seeder species—that is, a species that does not resprout but, rather, recruits profusely only after fire. As Jepson stated,

“ After fires, it reappears promptly on ‘burns,’ and fruits at the age of 5 or 6 years. It thus adapts itself to short fire intervals and is a true fire-type shrub ” (Jepson 1922 ).

Jepson also noted the large burls at the base of some Arctostaphylos species and the important role they played in postfire resprouting (figure  1 ). This is a significant observation as other early botanists who studied such structures were perplexed with their functional significance and did not draw this conclusion. For example, Kerr ( 1925 ) wrote an extensive paper on these structures that develop in seedlings of many Eucalyptus species and named them lignotubers , a structure today widely viewed as a fire adaptive trait (Keeley and Pausas 2022 ). He noted that, after fire, resprouts emerged from these lignotubers, but he attached no great significance to the role of fire in the evolution of this trait.

Two of the manzanitas from Jepson (1916): Arctostaphylos glandulosa (left) and Arctostaphylos nummularia (right). The former is an obligate resprouter (note the basal burl) and the latter is an obligate seeder (note the even-aged cohort). Manzanitas were among the first plants that made scientists think about the role of fire in plant evolution (Jepson 1922).

Two of the manzanitas from Jepson ( 1916 ): Arctostaphylos glandulosa (left) and Arctostaphylos nummularia (right). The former is an obligate resprouter (note the basal burl) and the latter is an obligate seeder (note the even-aged cohort). Manzanitas were among the first plants that made scientists think about the role of fire in plant evolution (Jepson 1922 ).

A few years later, in 1926, Walter W. Hough (1859–1935) implicitly suggested fire adaptations in trees subjected to recurrent fires (thick bark), although he did not mention the term adaptation :

“ If, as appears probable, forests have been swept by fire at intervals throughout their history, it is likely that there has been established in some tree species a resistance to the effect of heat. There may be seen in the thickening of the bark near the ground perhaps a protective device ” (Hough 1926 : 8).

In 1958, Tom M. Harris (1903–1983) published a paper on Mesozoic (Jurassic) fires, perhaps the first paper providing evidence (fossil charcoal) of fires in deep time. This evidence of ancient fires was largely ignored until quite recently. He also mentioned the possibility of species (animals and plants) from open ecosystems having evolved because of gaps opened by fire:

“ The objection usually urged against accepting fusian as charcoal produced by fire is that there is too much of it and in too many layers.… If occurring at intervals of a few centuries they wildfires might not change the forest climax, but they would destroy patches of young or mature forests and provide a home for the animals and plants of open ground or younger stages in the forest succession.… It would help us to understand the origin of the vast number of species which today seem to depend on fire; they may have already evolved in strength and have been ready to seize the increase opportunities offered by man ” (Harris 1958 : 449, 453).

Inspired by Jepson's early observations, Philip V. Wells (1928–2004), also referring to postfire nonresprouting seeders of the genus Arctostaphylos and Ceanothus (California), concluded,

“ A quickening of the tempo of evolution in major sections of Arctostaphylos and Ceanothus in California would appear to stem from their unique abandonment of the conservative, crown-sprouting mode of reproduction in favor of a nonsprouting, obligately seeding response to recurrent fire that results in a greater frequency and intensity of selection ” (Wells 1969 : 266)

And in 1970, Robert W. Mutch (born 1934) proposed his famous hypothesis on the evolution of flammability (see Pausas et al. 2017 for a current review on this topic):

“ Fire-dependent plant communities burn more readily than non-fire-dependent communities because natural selection has favored development of characteristics that make them more flammable ” (Mutch 1970 ).

During the 1970s, many other researchers began to accept the evolutionary role of fire, but it was not yet widely recognized. For instance, Axelrod wrote extensively on the origin of Californian flora (Axelrod 1973 , Raven and Axelrod 1978 , 1989 ) and assigned no role to fire in the evolution of this fire-adapted flora (Keeley et al. 2012 ).

It is interesting to note that, outside the English-speaking world, there were also some researchers aware of the evolutionary role of fire during the 1970s. This is the case of Leopoldo M. Coutinho (1934–2016, Sao Paulo, Brazil), who discussed fire adaptations in South American savannas during that time (e.g., Coutinho 1976 , 1977 ); his view was quite advanced in understanding the evolutionary role of fire compared with the mainstream knowledge (Pausas 2017 ).

Since the 1970s, there have been a large number of studies showing the adaptive value of many plant traits to fire (table  1 ). Many of those studies focused on Mediterranean-type vegetation and pine or oak woodlands, following Jepson, Wells, and Mutch, and only more recently, evolutionary fire ecology has been applied to savannas (e.g., Simon and Pennington 2012 , Maurin et al. 2014 ). At the same time, there was the recognition that fires were ancient at the geological scale (for a review, see Scott 2018 ), and, therefore, rescuing Harris’s ideas. The availability of molecular phylogenies in the 2000s boosted the study of plant evolution; ancestral reconstructions along phylogenies showed that fire adaptive traits are very old and their origin are often linked to geological times when fire activity was especially high (Bytebier et al. 2011 , Crisp et al. 2011 , He et al. 2012 , Crisp and Cook 2013 , Lamont et al. 2019 ). Therefore, the existence of fire adaptations (traits shaped by fire), in addition to fire adaptive traits (traits that confer adaptive value under fires), is now unambiguous.

Summary of traits that are adaptive to fire-prone environments with some key references. Some of them are indicated as poorly known, but listed to stimulate further research.

SurvivalResprouting traits (plant survival)Root crownLignotuberWidespreadCrown firesPausas et al. ( )Paula et al. ( ), Pausas et al. ( )
Woody rhizomesSurface (and crown) firesMaurin et al. ( ), Pausas et al. ( )
Epicormic resproutingSurface fires (savannas), crown fires (e.g., eucalypts)Charles-Dominique et al. ( ), Pausas and Keeley ( )
Sunken stem budsEucalypts, savanna treesBurrows ( ), ( ), Charles-Dominique et al. ( )
Smoke-induced nutrient translocationPoorly knownRabideau-Childers et al. ( )
Stem survivalThick (outer) barkSurface firesJackson et al. ( ), Pausas ( )
Reduced flammabilitySurface fires, pines (self-pruning), savannas (corky strategy)Keeley and Zedler ( ), Dantas and Pausas ( ), Pausas et al. ( )
Reproduction and recruitmentHeat-released dormancyCrown fires (e.g., many Fabaceae, Cistaceae)Keeley ( ), Keeley and Fotheringham ( ), Pausas and Lamont ( )
Smoke-released dormancyCrown fires (e.g., many Ericaceae and Labiatae)Keeley ( ), Keeley and Fotheringham ( ), Pausas and Lamont ( ), Lamont and Pausas ( )
Seed traits enhancing seed survivalPoorly studiedGómez-González et al. ( ), Liyanage and Ooi ( ), Tangney et al. ( )
SerotinyCrown fires; large shrubs and treesLamont et al. ( )
Fire-stimulated floweringCrown and surface fires; mainly in geophytesLamont and Downes ( )
Increased flammability (chemically or structurally)Crown fires; associated to heat-released dormancy or serotinyBond and Midgley ( ), Pausas et al. ( )
Precocity (i.e., early reproduction)In woody nonresprouters under short fire intervalsGuiote and Pausas ( )
Elaiosomes (ant dispersal)Seed burial for insulation; Crown fires; poorly knownBerg ( ), Pausas and Lamont ( )
SurvivalResprouting traits (plant survival)Root crownLignotuberWidespreadCrown firesPausas et al. ( )Paula et al. ( ), Pausas et al. ( )
Woody rhizomesSurface (and crown) firesMaurin et al. ( ), Pausas et al. ( )
Epicormic resproutingSurface fires (savannas), crown fires (e.g., eucalypts)Charles-Dominique et al. ( ), Pausas and Keeley ( )
Sunken stem budsEucalypts, savanna treesBurrows ( ), ( ), Charles-Dominique et al. ( )
Smoke-induced nutrient translocationPoorly knownRabideau-Childers et al. ( )
Stem survivalThick (outer) barkSurface firesJackson et al. ( ), Pausas ( )
Reduced flammabilitySurface fires, pines (self-pruning), savannas (corky strategy)Keeley and Zedler ( ), Dantas and Pausas ( ), Pausas et al. ( )
Reproduction and recruitmentHeat-released dormancyCrown fires (e.g., many Fabaceae, Cistaceae)Keeley ( ), Keeley and Fotheringham ( ), Pausas and Lamont ( )
Smoke-released dormancyCrown fires (e.g., many Ericaceae and Labiatae)Keeley ( ), Keeley and Fotheringham ( ), Pausas and Lamont ( ), Lamont and Pausas ( )
Seed traits enhancing seed survivalPoorly studiedGómez-González et al. ( ), Liyanage and Ooi ( ), Tangney et al. ( )
SerotinyCrown fires; large shrubs and treesLamont et al. ( )
Fire-stimulated floweringCrown and surface fires; mainly in geophytesLamont and Downes ( )
Increased flammability (chemically or structurally)Crown fires; associated to heat-released dormancy or serotinyBond and Midgley ( ), Pausas et al. ( )
Precocity (i.e., early reproduction)In woody nonresprouters under short fire intervalsGuiote and Pausas ( )
Elaiosomes (ant dispersal)Seed burial for insulation; Crown fires; poorly knownBerg ( ), Pausas and Lamont ( )

Despite the fact that, in the 1970s, there was already evidence of fire acting as an evolutionary process, the idea is only slowly percolating into mainstream ecology and evolutionary biology. In fact, evolutionary fire ecology is still rarely considered in most ecology textbooks. Many studies on plant ecology and evolution have either ignored fire or treated it as an incidental process without adequately considering the ecological and evolutionary feedback loops among fire, vegetation, climate, and geology. Even less studied is the interaction of fire with other evolutionary forces (e.g., drought, species interactions) in the evolution of plant traits. There are examples in the scientific literature of failures to consider fire as a possible explanation when studying ecological and evolutionary processes in fire-prone ecosystems. For instance, some of the plant mortality attributed to climate warming may be better explained by fire history (Schwilk and Keeley 2012 ); seed dormancy has often been suggested to be an adaptation to seasonal climates, but a more proximal explanation may be that it is a response to the fires occurring in those climates (Pausas et al. 2022 ); and the diversity and distribution of many species cannot be explained without considering fire (Pausas and Lamont 2018 ). This fire blindness is likely to be a legacy of the cultural bias by early European naturalists (Pausas and Bond 2019 ). However, few ecological processes are so disrupting as wildfires; they can produce large-scale disturbances consuming massive amounts of plants and affecting all organisms in a biome. But fire-prone ecosystems are hotspots of diversity, not despite fires but, at least in part, because of them (Cowling 1987 , He et al. 2019 ). Below, we list some key contributions that fire ecology has made to evolutionary biology.

First, fire has affected plant communities from the very origin of land plants, in the Silurian (Glasspool et al. 2004 ); since then, fire regimes have been fluctuating as a consequence of changes in vegetation, climate, herbivores, and atmospheric oxygen concentration (Scott 2018 ). Therefore, fire is among the earliest disturbance processes in plant communities and among the earliest potential evolutionary pressures in land plants.

Second, fires were recurrent and predictable enough to select fire adaptive traits and contributed to the diversification of lineages, at least since the Cretaceous (Crisp et al. 2011 , He et al. 2012 ) but probably earlier (Keeley and Pausas 2022 ). Therefore, fire has contributed to the evolution of many plant traits since early plant evolution.

Third, fire likely contributed to the spread and evolution of large lineages such as early angiosperms (Bond and Scott 2010 ) and C 4 grasses (Keeley and Rundel 2005 , Scheiter et al. 2012 ), and it is therefore responsible for the rise of species-rich savannas (Simon et al. 2009, Maurin et al. 2014 ).

Fourth, fire ecology provides examples of how different regimes select for radically different adaptations across species. For instance, recurrent fires with low or with high intensity selects contrasting traits in plants (e.g., thick bark versus serotinous cones in pines; Keeley and Zedler 1998 , Pausas 2015b ). Similarly, different fire frequencies also select quite different traits: high frequency (resprouting) and moderate fire frequency (postfire seeding and the loss of resprouting; Pausas and Keeley 2014 ). At the landscape scale, different fire regimes generate different evolutionary frameworks for a given environmental conditions (e.g., savannas versus forests; Pausas and Bond 2020 ).

Finally, fire ecology provides evidence that different selective regimes generate trait divergences among populations and therefore the evolution of fire related traits (Gómez-González et al. 2011 , Pausas et al. 2012, Hernández-Serrano et al. 2013 , Vandvik et al. 2014 , Guiote and Pausas 2023 , Keeley 2023 ). Fire, by opening vegetation gaps, provides opportunities for the evolution of many light-demanding shade-intolerant species (Bond 2019 , Pausas and Lamont 2022 ).

That is, we cannot understand the biodiversity of our planet without considering the key evolutionary role of fire (“ a world without fires is like a sphere without roundness—i.e., we cannot imagine it ”; Pausas and Keeley 2009 ).

The thinking about the evolutionary role of fire started over a century ago and accelerated only recently. There is still much research to do before we fully understand the role of fire in the evolution of species. It is imperative to continue performing basic natural history observations (e.g., Keeley 2023 ), especially in Africa, South America, and Asia (e.g., Pausas et al. 2021 ), because there may be some fire strategies currently unknown. In fact, we still do not know the response of many species to fire, which makes predictions under fire regime changes difficult. This knowledge gap is occurring not only in little studied areas (e.g., Asia, tropical ecosystems) but also in areas that were not subject to fires until recently, and fire ecology therefore has had a limited tradition (e.g., Central Europe). So, plant trait databases that include fire traits need to be expanded in the number of species and geographically. There is also a bias toward woody plants in most databases. Databases may also need to include the fire characteristics (e.g., intensity, season, fire interval, fire type), because responses may be quite different depending on them. Trait databases, together with the increasing availability of phylogenetic information would also allow us to make more accurate estimations of the origin and evolutionary pathways of fire traits and, therefore, to better reconstruct the evolutionary fire history of our planet. A limitation of the phylogenetic approach is that early extinctions makes inferring deep patterns of character evolution rather difficult. For instance, fires were common in the Carboniferous, but we have been unable to trace back fire adaptations to that period. Is the abundance of rhizomes in ferns a response to those fires? (Pausas et al. 2018 ).

Another relevant question is how does evolution inform our predictions about how novel fire regimes will affect species and ecosystems. We have evidence that many fire related plant traits can evolve relatively rapidly under different fire regimes (e.g., Gómez-González et al. 2011 , Pausas et al. 2012, Hernández-Serrano et al. 2013 , Vandvik et al. 2014 , Guiote and Pausas 2023 , Keeley 2023 ). What are the limits of those evolutionary changes in the framework of our current fire regime changes? The ecoevolutionary feedback loops are also little explored; that is, plants can modify fire regimes, and these modifications can feed back to the plants with evolutionary consequences (Pausas and Bond 2022 ).

There is still little understanding of the evolutionary response to fire in the animal kingdom, especially because their responses are often behavioral rather than morphological, and it is therefore more difficult to pinpoint fire as a selective factor (Pausas and Parr 2018 ). However, changes in color, not only in invertebrates (Forsman et al. 2011 ), is still a research area as those changes may improve camouflage and thermoregulation in postfire conditions. Another little explored research area is how fire affects biotic interactions (e.g., García et al. 2016 ) and the evolutionary consequences and vice versa—that is, how biotic interactions affect the evolution of plant traits in interaction with fire (e.g., Talluto and Benkman 2014 ). For instances, how does fire temporarily disassembles the network of plant–animal interactions, and what are the consequences on the fitness of the interacting species? And how and to what extent do postfire dynamics reassemble those interactions?

Sequencing techniques that provide wide genome coverage are quickly advancing as their prices are plummeting, and they are therefore becoming more accessible to ecologists. This should allow us to explore the genetic footprint of traits selected by fire, as well as to study fire-driven population divergence and field heritability of fire traits (e.g., Castellanos et al. 2015 ). Epigenetics is also quickly advancing, and chances are that some of the fire products (smoke, ash) could produce some adaptive epigenetic changes; even processes such as postfire resprouting are likely to affect the epigenetic mosaicism with fitness consequences (Herrera et al. 2022 ). Therefore, epigeneitcs may be a fruitful research area for understanding quick postfire changes. Collectively, these advances should contribute to showing that fire has played an important evolutionary role.

This study was undertaken within the framework of the FocScales project (PROMETEO/2021/040, Generalitat Valenciana). Any use of trade, product, or firm names is for descriptive purposes only and does not imply endorsement by the US government.

Juli G. Pausas (Conceptualization, Formal analysis, Investigation, Methodology, Project administration,Visualization, Writing – original draft), and Jon E. Keeley (Writing – review & editing).

Author Biography

Juli G. Pausas ( [email protected] , [email protected] ) is affiliated with the Centro de Investigaciones sobre Desertificación, Consejo Superior de Investigaciones Cientificas, in Montcada, Spain. Jon E. Keeley is affiliated with the Sequoia–Kings Canyon Field Station, at theWestern Ecological Research Center, US Geological Survey, in Three Rivers, California, and with the Department of Ecology and Evolutionary Biology at the University of California–Los Angeles, in Los Angeles, California, in the United States

Álvarez-Ruiz L , Belliure J , Pausas JG. 2021 . Fire-driven behavioral response to smoke in a Mediterranean lizard . Behavioral Ecology 32 : 662 – 667 .

Google Scholar

Axelrod DI. 1973 . History of the Mediterranean ecosystem in California . Pages 225 – 277 in di Castri F , Mooney HA , eds. Mediterranean Ecosystems: Origin and Structure . Springer .

Google Preview

Axelrod DI. 1989 . Age and origin of chaparral . Pages 7 – 19 in Keeley SC , ed. The California Chaparral: Paradigms Reexamined. Science Series no. 34 . Natural History Museum of Los Angeles County .

Berg RY. 1975 . Myrmecochorous plants in Australia and their dispersal by ants . Australian Journal of Botany 23 : 475 – 508 .

Bond WJ. 2019 . Open Ecosystems: Ecology and Evolution Beyond the Forest Edge . Oxford University Press .

Bond WJ , Midgley JJ. 1995 . Kill thy neighbour: An individualistic argument for the evolution of flammability . Oikos 73 : 79 – 85 .

Bond WJ , Scott AC. 2010 . Fire and the spread of flowering plants in the Cretaceous . New Phytologist 188 : 1137 – 1150 .

Burrows GE. 2002 . Epicormic strand structure in Angophora, Eucalyptus and Lophostemon (Myrtaceae): Implications for fire resistance and recovery . New Phytologist 153 : 111 – 131 .

Burrows GE. 2013 . Buds, bushfires, and resprouting in the eucalypts . Australian Journal of Botany 61 : 331 – 349 .

Bytebier B , Antonelli A , Bellstedt DU , Linder HP. 2011 . Estimating the age of fire in the Cape flora of South Africa from an orchid phylogeny . Proceedings of the Royal Society B 278 : 188 – 195 .

Castellanos MC , González-Martínez SC , Pausas JG . 2015 . Field heritability of a plant adaptation to fire in heterogeneous landscapes . Molecular Ecology 24 : 5633 – 5642 .

Charles-Dominique T , Beckett H , Midgley GF , Bond WJ. 2015 . Bud protection: A key trait for species sorting in a forest–savanna mosaic . New Phytologist 207 : 1052 – 1060 .

Coutinho LM. 1976 . Contribuição ao Conhecimento do Papel Ecológico das Queímadas na Floraçao de Especies do Cerrado . Universidade de Sao Paulo .

Coutinho LM. 1977 . Aspectos ecológicos do fogo no Cerrado: II. As queimadas ea dispersao de sementes . Boletim de Botânica 5 : 57 – 64 .

Cowling RM. 1987 . Fire and its role in coexistence and speciation in Gondwanan shrublands . South African Journal of Science 83 : 106 – 112 .

Crisp MD , Cook LG. 2013 . How was the Australian flora assembled over the last 65 million years? A molecular phylogenetic perspective . Annual Review of Ecology, Evolution, and Systematics 44 : 303 – 324 .

Crisp MD , Burrows GE , Cook LG , Thornhill AH , Bowman DMJS. 2011 . Flammable biomes dominated by eucalypts originated at the Cretaceous–Palaeogene boundary . Nature Communications 2 : 193 .

Dantas VL , Pausas JG. 2013 . The lanky and the corky: Fire-escape strategies in savanna woody species . Journal of Ecology 101 : 1265 – 1272 .

Dobzhansky T . 1973 . Nothing in Biology Makes Sense except in the Light of Evolution . The American Biology Teacher 35 : 125 – 129 .

Forsman A , Karlsson M , Wennersten L , Johansson J , Karpestam E. 2011 . Rapid evolution of fire melanism in replicated populations of pygmy grasshoppers . Evolution 65 : 2530 – 2540 .

García Y , Castellanos MC , Pausas JG. 2016 . Fires can benefit plants by disrupting antagonistic interactions . Oecologia 182 : 1165 – 1173 .

Glasspool IJ , Edwards D , Axe L. 2004 . Charcoal in the Silurian as evidence for the earliest wildfire . Geology 32 : 381 – 383 .

Gómez-González S , Torres-Díaz C , Bustos-Schindler C , Gianoli E. 2011 . Anthropogenic fire drives the evolution of seed traits . Proceedings of the National Academy of Sciences 108 : 18743 – 18747 .

Guiote C , Pausas JG. 2023 . Fire favors sexual precocity in a Mediterranean pine . Oikos 2023 : e09373 .

Guthrie RD. 1967 . Fire melanism among mammals . American Midland Naturalist 77 : 227 – 230 .

Harris TM. 1958 . Forest Fire in the Mesozoic . Journal of Ecology 46 : 447 – 453 .

He T , Pausas JG , Belcher CM , Schwilk DW , Lamont BB. 2012 . Fire-adapted traits of Pinus arose in the fiery Cretaceous . New Phytologist 194 : 751 – 759 .

He T , Lamont BB , Pausas JG. 2019 . Fire as a key driver of Earth's biodiversity . Biological Reviews 94 : 1983 – 2010 .

Hernández-Serrano A , Verdú M , González-Martínez SC , Pausas JG . 2013 . Fire structures pine serotiny at different scales . American Journal of Botany 100 : 2349 – 2356 .

Herrera CM , Medrano M , Bazaga P , Alonso C . 2022 . Ecological significance of intraplant variation: Epigenetic mosaicism in Lavandula latifolia plants predicts extant and transgenerational variability of fecundity-related traits . Journal of Ecology 110 : 2555 – 2567 .

Hough W. 1926 . Fire as an Agent in Human Culture . US Government Printing Office .

Hutto RL. 2008 . The ecological importance of severe wildfires: Some like it hot . Ecological Applications 18 : 1827 – 1834 .

Jackson JF , Adams DC , Jackson UB. 1999 . Allometry of constitutive defence: A model and a comparative test with tree bark and fire regime . American Naturalist 153 : 614 – 632 .

Jepson WL. 1916 . Regeneration in manzanita . Madroño 1 : 3 – 12 .

Jepson WL. 1922 . Revision of the California species of the genus Arctostaphylos . Madroño 1 : 76 – 86 .

Keeley JE. 1991 . Seed germination and life history syndromes in the California chaparral . Botanical Review 57 : 81 – 116 .

Keeley JE. 2023 . Spatial and temporal strategies of resprouting and seeding in a chaparral shrub species . Ecology 104 : e3984 .

Keeley JE , Fotheringham CJ. 2000 . Role of fire in regeneration from seeds . Pages 311 – 330 in Fenner M , ed. Seeds: The Ecology of Regeneration in Plant Communities , 2nd ed. CAB International .

Keeley JE , Pausas JG. 2022 . Evolutionary ecology of fire . Annual Review of Ecology, Evolution, and Systematics 53 : 203 – 225 .

Keeley JE , Rundel PW. 2005 . Fire and the Miocene expansion of C4 grasslands . Ecology Letters 8 : 683 – 690 .

Keeley JE , Zedler PH. 1998 . Evolution of life histories in Pinus . Pages 219 – 250 in Richardson DM , ed. Ecology and Biogeography of Pinus . Cambridge University Press .

Keeley JE , Pausas JG , Rundel PW , Bond WJ , Bradstock RA. 2011 . Fire as an evolutionary pressure shaping plant traits . Trends in Plant Science 16 : 406 – 411 .

Keeley JE , Bond WJ , Bradstock RA , Pausas JG , Rundel PW. 2012 . Fire in Mediterranean Ecosystems: Ecology, Evolution, and Management . Cambridge University Press .

Kerr LR. 1925 . The lignotubers of eucalypt seedlings . Proceedings of the Royal Society of Victoria 37 : 79 – 97 .

Keynes RD. 2001 . Charles Darwin's Beagle Diary . Cambridge University Press .

Komarek EV. 1969 . Fire and animal behavior . Pages 160–207 in Komarek EV, ed. Proceedings Tall Timbers Fire Ecology Conference, vol 9: Tall Timbers Fire Ecology Conference . Tall Timbers Research Station .

Lamont BB , Downes KS. 2011 . Fire-stimulated flowering among resprouters and geophytes in Australia and South Africa . Plant Ecology 212 : 2111 – 2125 .

Lamont BB , Pausas JG. 2023 . Seed dormancy revisited: Dormancy-release pathways and environmental interactions . Functional Ecology 37 : 1106 – 1125 .

Lamont BB , He T , Yan Z . 2019 . Evolutionary history of fire-stimulated resprouting, flowering, seed release and germination . Biological Reviews 94 : 903 – 928 .

Lamont BB , Pausas JG , He T , Witkowski ETF , Hanley ME. 2020 . Fire as a selective agent for both serotiny and nonserotiny over space and time . Critical Reviews in Plant Sciences 39 : 140 – 172 .

Lillywhite HB , Friedman G , Ford N. 1977 . Color matching and perch selection by lizards in recently burned chaparral . Copeia 1977 : 115 – 121 .

Liyanage GS , Ooi MKJ. 2018 . Seed size-mediated dormancy thresholds: A case for the selective pressure of fire on physically dormant species . Biological Journal of the Linnean Society 123 : 135 – 143 .

Maurin O , Davies TJ , Burrows JE , Daru BH , Yessoufou K , Muasya AM , van der Bank M , Bond WJ. 2014 . Savanna fire and the origins of the ‘underground forests’ of Africa . New Phytologist 204 : 201 – 214 .

Mutch RW. 1970 . Wildland fires and ecosystems: A hypothesis . Ecology 51 : 1046 – 1051 .

Nicholas FW , Nicholas J. 2008 . Charles Darwin in Australia . Cambridge University Press .

Nimmo DG , Carthey AJR , Jolly CJ , Blumstein DT. 2021 . Welcome to the pyrocene: Animal survival in the age of megafire . Global Change Biology 27 : 5684 – 5693 .

Nowack J , Stawski C , Körtner G , Geiser F. 2018 . Physiological and behavioral responses of an arboreal mammal to smoke and charcoal-ash substrate . Physiology and Behavior 184 : 116 – 121 .

Paula S , Naulin PI , Arce C , Galaz C , Pausas JG. 2016 . Lignotubers in Mediterranean basin plants . Plant Ecology 217 : 661 – 676 .

Pausas JG. 2015a . Bark thickness and fire regime . Functional Ecology 29 : 315 – 327 .

Pausas JG. 2015b . Evolutionary fire ecology: Lessons learned from pines . Trends in Plant Science 20 : 318 – 324 .

Pausas JG. 2017 . Homage to L. M. Coutinho: Fire adaptations in Cerrado plants . International Journal of Wildland Fire 26 : 249 – 251 .

Pausas JG , Bond WJ. 2019 . Humboldt and the reinvention of nature . Journal of Ecology 107 : 1031 – 1037 .

Pausas JG , Bond WJ. 2020 . Alternative biome states in terrestrial ecosystems . Trends in Plant Science 25 : 250 – 263 .

Pausas JG , Bond WJ. 2022 . Feedbacks in ecology and evolution . Trends in Ecology and Evolution 37 : 637 – 644 .

Pausas JG , Keeley JE. 2009 . A burning story: The role of fire in the history of life . Bioscience 59 : 593 – 601 .

Pausas JG , Keeley JE. 2014 . Evolutionary ecology of resprouting and seeding in fire-prone ecosystems . New Phytologist 204 : 55 – 65 .

Pausas JG , Keeley JE. 2017 . Epicormic resprouting in fire-prone ecosystems . Trends in Plant Science 22 : 1008 – 1015 .

Pausas JG , Lamont BB. 2018 . Ecology and biogeography in 3D: The case of the Australian Proteaceae . Journal of Biogeography 45 : 1469 – 1477 .

Pausas JG , Lamont BB. 2022 . Fire-released seed dormancy: A global synthesis . Biological Reviews 97 : 1612 – 1639 .

Pausas JG , Parr CL. 2018 . Towards an understanding of the evolutionary role of fire in animals . Evolutionary Ecology 32 : 113 – 125 .

Pausas JG , Keeley JE , Schwilk DW. 2017 . Flammability as an ecological and evolutionary driver . Journal of Ecology 105 : 289 – 297 .

Pausas JG , Lamont BB , Paula S , Appezzato-da-Glória B , Fidelis A. 2018 . Unearthing belowground bud banks in fire-prone ecosystems . New Phytologist 217 : 1435 – 1448 .

Pausas JG , Su W-H , Luo C , Shen Z. 2021 . A shrubby resprouting pine with serotinous cones endemic to Southwest China . Ecology 102 : e03282 .

Pausas JG , Lamont BB , Keeley JE , Bond WJ. 2022 . Bet-hedging and best-bet strategies shape seed dormancy . New Phytologist 236 : 1232 – 1236 .

Poulton EB. 1915 . The habits of the Australian buprestid “fire-beetle” Merimna atrata , Lap. Et Gory . Proceedings of the Entomological Society of London 1 : iii – iv .

Poulton EB. 1926 . Protective resemblance borne by certain African insects to the blackened areas caused by grass fires. Verhandlungen des IIIen internationales entomologisches . Kongress Zürich 1925 2 : 12 – 94 .

Rabideau-Childers R et al.  2022 . Evidence of nutrient translocation in response to smoke exposure by the East African ant acacia, Vachellia drepanolobium . Ecology and Evolution 12 : e8244 .

Raven PH , Axelrod DI. 1978 . Origin and Relationships of the California Flora . University of California Press .

Scheiter S , Higgins S I , Osborne C P , Bradshaw C , Lunt D , Ripley B S , Taylor L L , Beerling D J . 2012 . Fire and fire-adapted vegetation promoted C4 expansion in the late Miocene . New Phytologist 195 : 653 – 666 .

Schwilk DW , Keeley JE. 2012 . A plant distribution shift: Temperature, drought or past disturbance? PLOS ONE 7 : e31173 .

Scott AC. 2018 . Burning Planet: The Story of Fire through Time . Oxford Univeristy Press .

Simon MF , Pennington T. 2012 . Evidence for adaptation to fire regimes in the tropical savannas of the Brazilian Cerrado . International Journal of Plant Sciences 173 : 711 – 723 .

Stawski C , Körtner G , Nowack J , Geiser F. 2015 . The importance of mammalian torpor for survival in a post-fire landscape . Biology Letters 11 : 20150134 .

Stromberg M. 1997 . Taricha torosa (California newt) response to fire . Herpetological Review 28 : 82 – 84 .

Talluto MV , Benkman CW. 2014 . Conflicting selection from fire and seed predation drives fine-scaled phenotypic variation in a widespread North American conifer . Proceedings of the National Academy of Sciences 111 : 9543 – 9548 .

Tangney R , Merritt DJ , Fontaine JB , Miller BP. 2019 . Seed moisture content as a primary trait regulating the lethal temperature thresholds of seeds . Journal of Ecology 107 : 1093 – 1105 .

Vandvik V , Töpper JP , Cook Z , Daws MI , Heegaard E , Måren IE , Velle LG. 2014 . Management-driven evolution in a domesticated ecosystem . Biology Letters 10 : 20131082 .

Wells PV. 1969 . The relation between mode of reproduction and extent of speciation in woody genera of the California chaparral . Evolution 23 : 264 – 267 .

Month: Total Views:
August 2023 802
September 2023 424
October 2023 192
November 2023 200
December 2023 490
January 2024 191
February 2024 150
March 2024 184
April 2024 123
May 2024 120
June 2024 94
July 2024 92
August 2024 90

Email alerts

Citing articles via.

  • Recommend to your Library

Affiliations

  • Online ISSN 1525-3244
  • Copyright © 2024 American Institute of Biological Sciences
  • About Oxford Academic
  • Publish journals with us
  • University press partners
  • What we publish
  • New features  
  • Open access
  • Institutional account management
  • Rights and permissions
  • Get help with access
  • Accessibility
  • Advertising
  • Media enquiries
  • Oxford University Press
  • Oxford Languages
  • University of Oxford

Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide

  • Copyright © 2024 Oxford University Press
  • Cookie settings
  • Cookie policy
  • Privacy policy
  • Legal notice

This Feature Is Available To Subscribers Only

Sign In or Create an Account

This PDF is available to Subscribers Only

For full access to this pdf, sign in to an existing account, or purchase an annual subscription.

What is fire ecology?

Fire is one of the major shaping influences of global ecosystems, and most ecosystems have a very interesting history with fire, one that has changed a great deal in the modern era.

Fire ecology is the study of fires and fire regimes in global forest, prairie, shrubland, chaparral, meadow, and savannah ecosystems. Fire ecologists study how changes in regime factors, like the frequency, severity, and extent of fires can provide favorable or damaging influences on ecosystems. They also study the interactions between fires and other ecological patterns and processes; how they work together or in conflict, to vary outcomes on the land. 

Many global ecosystems have evolved with fire as an essential process that creates and renews habitats. In fire prone ecosystems, native plants and animals have evolved an essential relationship with fire influencing not only habitat conditions but key aspects of their life history. Fire ecology practitioners work to understand these relations and, often by means of intentional burning practices and other methods, restore these intimate relationships with fire. 

Introduction to Fire ecology

World of Wildland Fire Videos : 10 videos that provide an overview of fire ecology and managment topics.

Fire Ecology : A chapter by AFE Board Member Dr. Robert Keane in the book, Encyclopedia of Wildfires and Wildland-Urban Interface (WUI) Fires.

Intro to Fire Ecology Across the US : An introductory chapter in the 2021 book, Fire Ecology and Management: Past, Present, and Future of US Forested Ecosystems. Managing Forest. Several chapters of this book are available online through the USFS TreeSearch and focus on fire ecology in specific regions of the US.

Fire Ecology 2.0 : A plenary talk at the 2017 Fire Congres by Dr. Leda Kobziar that focuses on the development and evolution of the fire ecology field.

fire ecology research

Fire Ecology Journal : Peer-reviewed scientific articles that are available for free online.

Fire Lab Seminar Series, USFS Missoula Fire Science Lab : Seminars presented by Fire Lab employees and other researchers from throughout the world that cover current wildland fire research and management topics. Several recorded seminars are available.

Joint Fire Science Program: Reports and findings from JFSP-funded research projects are available on their website. The JSFP Fire Science Exchange Network is an amazing resource for location specific publications, workshops, and field tours.

Program and Degrees

Many colleges and universities offer wildland fire courses. AFE has an academic certification program to recognize programs which prepare future fire ecology and management professionals. Click here for a list of certified programs .

Curriculum Resources for K-12 Education

FireWorks: Interactive, hands-on materials to study wildland fire. FireWorks has specialized curricula for several regions of the U.S. as well as a generic curricula that can be adapted to other areas.

Fire Ecology Learning Lab: Curriculum developed by the Southwest Fire Science Consortium to inspire our future natural resource stewards and community leaders, with materials for both teachers and non-formal educators.

Southern Oregon Fire Ecology Education: Standards-aligned, trauma-informed curriculum for grades K-12 that uses fire as a lens for STEAM learning, land management literacy, and entry into fire-related career pathways.​ Available in English and Spanish, for online and classroom learning.

FECO cover image.jpg

AFE’s journal,  Fire Ecology , publishes peer-reviewed articles on all ecological and management aspects relating to wildland fire. Freely access journal articles and get details about submitting a manuscript.

  • Skip to global NPS navigation
  • Skip to the main content
  • Skip to the footer section

research paper on fire ecology

Exiting nps.gov

Fire ecology and research.

Scientific information is the cornerstone of safe and successful fire management.

The field of fire ecology seeks to understand when fires occurred in the past, how plants and animals in various environments respond and adapt to fire, and how fires and their effects may change in the future.

National Park Service fire managers strive to ensure the most current science-based information is integrated into fire and land management goals, decisions, and practices. This work includes collaborating with resource managers and scientists to develop fire management objectives that meet land management goals, designing and implementing monitoring programs to determine if objectives are met, and identifying questions that need to be answered through research studies. Providing this type of information to managers is critical to ensuring a scientifically based fire management program that will continue to improve as new knowledge is gained.

The fire effects monitoring program allows park managers to document basic information, to detect trends, and to ensure that each park meets its fire and resource management objectives.

Learn about Fire Research & Fire Ecology

Fire managers look to scientific research for guidance on how prescribed burns and lightning-started fires can can be used to benefit ecological systems. They examine historic fire regimes and past cultural practices, and use predictive modeling to study future fire effects under different climates and communities. Fire research is also important for forecasting fire behavior and spread across the landscape for suppression tactics, planning fuels reduction treatments, and fire season preparedness.

National park units host many scientific research studies. Parks are highly valued as study sites because the land has been protected, in some cases for more than 100 years. Research projects may be funded by a variety of sources both internal and external to the agency. In 1998, the Joint Fire Science Program , a partnership of six federal wildland, fire, and research organizations, was established to provide scientific information and support for fuel and fire management programs.

FFI (FEAT/FIREMON Integrated) FFI is a monitoring software tool designed to assist managers with collection, storage, and analysis of ecological information. This database management system was developed to support immediate and long-term monitoring and reporting of fire effects, and its use encourages cooperative, interagency information sharing. Fire Monitoring Handbook (PDF) The fire monitoring methods described in the NPS Fire Monitoring Handbook allow park managers to document basic information, detect trends, and ensure that each park meets its fire and resource management objectives. From identified trends, park staff can articulate concerns, develop hypotheses, and identify specific research studies to develop solutions to problems.

Monitoring Trends in Burn Severity The Monitoring Trends in Burn Severity project addresses the need to quantify fire effects over large, often-remote regions and long time intervals. It reflects collaborative efforts to bring previous research into operational implementation for fire managers and scientists.

The Composite Burn Index (CBI) Photo Series The Composite Burn Index (CBI) Photo Series uses CBI plot data and photos to illustrate the range of burn severity encountered in different ecosystems of the United States.

Measuring and Monitoring Plant Populations (5.12 MB) Interagency technical monitoring reference.

Last updated: April 2, 2020

U.S. flag

An official website of the United States government

Here's how you know

Official websites use .gov A .gov website belongs to an official government organization in the United States.

Secure .gov websites use HTTPS A lock ( Lock A locked padlock ) or https:// means you’ve safely connected to the .gov website. Share sensitive information only on official, secure websites.

  • Data and Tools
  • National Forest Service Library

Introduction to fire ecology across USA forested ecosystems: Past, present, and future

Authors: , Beverly S. Collins, , Michael C. Stambaugh, Gary R. Wein
Year: 2021
Type: Book Chapter
Station: Southern Research Station
Source: In: Greenberg, Cathryn H.; Collins, Beverly., eds. Fire Ecology and Management: Past, Present, and Future of US Forested Ecosystems. Managing Forest Ecosystems 39/Springer
  • Request Info
  • Faculty & Staff
  • Job Seekers
  • Scholarships & Aid
  • Student Life

Jacques Rinchard ( [email protected] )

  • Daily Eagle

Environmental Science & Ecology Alum & Faculty Publish Peer-Reviewed Paper

Jarrod Ludwig (BS/MS’24) and Dr. Jacques Rinchard published a paper in Journal of Great Lakes Research.

This paper, based on Jarrod’s SURP project from his undergraduate studies in 2022, provides insights into the reproductive spawning strategy of the deepwater sculpin through histological analysis.

Ludwig, J., Weidel, B., O’Malley, B., Connerton, M., and Rinchard, J., 2024. Histological analysis of deepwater sculpin ovaries supports single spawning reproductive strategy. Journal of Great Lakes Research , 50, 102375.

Jarrod is currently working as an Aquatic Biologist at the NYS Department of Environmental Conservation in Dunkirk, NY.

  • More Information

Wildfires and Real Estate Values in California

research paper on fire ecology

Leila Bengali

research paper on fire ecology

Fernanda Nechio

Stephanie A. Stewart

Download PDF (647 KB)

FRBSF Economic Letter 2024-22 | August 26, 2024

Wildfires have been a concern in California for decades. The intensity of these events has increased recently, with particularly large and destructive fire seasons between 2018 and 2021. Analysis shows that distance from high fire-risk zones had little impact on residential housing values in the past. However, that has changed since the late 2010s, coinciding with more extensive fire damage to land and structures across the state. Insurance availability appears to help little in preserving home values in areas that are considered more at risk.

Wildfires have damaged property in the state of California for decades, and fire risk in the state is elevated relative to most of the United States (Aylward and Oliveira 2020). Historical data from California’s state fire agency, CAL FIRE, recorded over 300,000 fire episodes in the state between 1987 and 2022. The size and intensity of these episodes have increased in recent years, along with their estimated costs. In the 1990s, a little over 415,000 acres burned on average each year; this annual average increased to just over 775,000 acres in the 2010s. At the same time, the number of structures destroyed by fires increased from about 355 yearly on average in the 1990s to an annual average of about 4,055 in the 2010s. This increase is due partly to a series of especially damaging fires in 2017 and 2018 and partly to an increase in residential use of areas deemed as high fire-risk zones (Mockrin et al. 2023).

The increases in severity, structure damage, and residential use of areas with high fire risk suggest that wildfire risk could negatively affect the residential real estate market in California, particularly as the state continues to struggle with a shortage of available housing units. Indeed, earlier research indicates that wildfires negatively impact residential real estate values of properties located near but generally not inside burned areas in southern California (Mueller, Loomis, and González-Cabán 2009).

In this Economic Letter , we estimate the effects of wildfire risk as measured by the distance from recent wildfires on residential real estate values. Our results suggest that property values have been more adversely impacted in recent years by being close to past wildfires than was the case previously. Moreover, while having insurance can help mitigate some of the costs associated with fire episodes, our results suggest that insurance does little to improve the adverse effects on property values.

Wildfires and residential areas in recent decades

Figure 1 summarizes wildfire activity in California from 1984 through 2021. It shows the number of wildfires and the average acreage burned per fire for wildfires exceeding 1,000 acres. The figure shows that, while the number of wildfires each year does not follow any particular trend over time, the area burned by wildfires has increased substantially, particularly in recent years.

Figure 1 California wildfires and acres burned

research paper on fire ecology

At the same time, the number of homes built in areas deemed as high risk has also increased over time. For example, a 2023 report from the USDA Forest Service estimated that the share of housing in the wildland-urban interface increased about 40% in California from 1990 to 2020 (Mockrin et al. 2023). The wildland-urban interface is generally regarded as having high fire risk.

Wildfire risk and housing values

The increasing intensity of wildfires and the growing exposure to high-risk areas have implications for real estate markets in California. We examine this issue using data from different sources. We use wildfire data from the Monitoring Trends in Burn Severity (MTBS) database of 1,000-plus acre fires, particularly information about each fire’s burn perimeter and ignition date, to measure past fire activity. Housing market data come from annual parcel-level administrative tax records obtained through CoreLogic. These data contain information on each property’s value, characteristics such as lot and building size, and location including zip code. We focus on single-family owner-occupied homes and analyze annual data at the zip code level. To measure wildfire risk, we identify the five wildfires closest to each zip code in each year, calculate the distances between the zip code and these five wildfire burn perimeters, and take an average. We use the geographic center of each zip code to calculate distances to the fire perimeters.

We estimate the relationship between distance from past wildfires and residential real estate values, controlling for other factors that can help explain variation in home values over time and across zip codes. These factors include property characteristics such as lot size, building square footage, and other property amenities. Importantly, we also account for trends over time by including average home values in the state and for typical differences in home values between zip codes by using zip code level averages. This means that the relationships we identify between wildfires and home values are driven by comparisons within a zip code rather than comparisons across zip codes. We omit zip codes with geographic centers within about 3 miles (5 kilometers) of fire perimeters to avoid using data on homes that may have been destroyed by fire. Finally, we estimate this relationship for wildfires that happened in the current year, the year before, two years before, and three years before to examine how a previous fire may affect values today. For brevity, here we focus on the effects of wildfires that happened three years before, since those patterns are similar to the patterns from earlier past fires.

Figure 2 reports the estimated relationship between the average distance in hundreds of miles to the zip code’s five closest wildfires and the average home values in the zip code. The blue dot shows the results using the first part of our sample, from 2008 to 2017. The green dot reports results for the most recent period, from 2018 to 2021. The red dot will be discussed in a later section. The bars around each data point show the statistical significance within a 95% confidence range. The positive values indicate that homes farther from past fires tend to have higher property values. Comparing the two results shows that the estimated relationship between distance from fire zones and home values was stronger in the more recent sample. While the change is notable, the effects are relatively small. Even in the late sample, being farther from past fires is associated with a boost in home value of about 2% for homes of average value.

Figure 2 Relationship between wildfire distance and home values

research paper on fire ecology

This change in patterns roughly aligns with the increasing wildfire intensity in California. The recent large fires may have changed homeowners’ perceptions of fire risk, which could alter how they view the tradeoff between amenities associated with living in risky areas and potential damages from wildfires (Donovan, Champ, and Butry 2007).

Using our estimates, we calculate the cumulative average effect of wildfires in 2021 and the three years prior on home values in each zip code. This calculation takes into account how far each zip code was from the closest five fires in 2021, 2020, 2019, and 2018. Figure 3 shows these estimates relative to the statewide average—that is, the difference between the estimated cumulative effect for each zip code and the average cumulative effect in the state.

Figure 3 Cumulative effects of 2018-2021 fires on home values

research paper on fire ecology

The figure shows wide variation across zip codes. In particular, coastal regions in central and northern California and arid desert regions in the extreme south experienced benefits relative to the average, shown as positive values, as these areas were farther than average from wildfires. In contrast, in vegetated and mountainous areas around Los Angeles and in the Sierras, wildfires lowered home values relative to the average, as these areas were closer to where wildfires burned.

Does insurance help home values?

Since our results indicate that wildfire risk may lower home values, we assess whether homeowners can counterbalance this risk with homeowner’s insurance.

Homeowners can obtain insurance through the private market or through the state-created California FAIR plan. The latter option is an alternative for homeowners who are not able to obtain insurance in the private market. The FAIR plan costs more and offers less generous coverage, protecting only the homeowner’s dwelling, as opposed to most plans that also cover personal belongings and have other benefits. As such, the FAIR plan use gives an indication of the quality of insurance homeowners can access. The FAIR plan market share is small for the residential market we study—about 3% in 2021—and its market share and market share growth vary across the state. Areas with higher FAIR plan use and growth tend to be those that face higher wildfire risk, such as hilly, mountainous, or heavily forested regions.

We use policy-level insurance data from the California Department of Insurance (CDI), which allows us to estimate insurance coverage rates for different types of policies. We add two measures of zip-code level insurance coverage to our model. The first is the percent of homes with private or public insurance. The second is a proxy for FAIR plan use, which captures coverage quality. Although the policy-level CDI data do not explicitly identify which plans are FAIR plans, the percent of homes with “dwelling only” insurance gives a good indication. Because of the noted association between FAIR plan use and wildfire risk, we also control for the proportion of insured homes that are designated as having high fire risk.

Using these insurance data, we revisit our analysis from Figure 2 to control for insurance access, shown by the red dot in Figure 2. Comparing the green and the red dots shows that controlling for risk classification and insurance access does little to limit the impact of distance to fire zones on home values.

This Letter assesses how living with wildfire risk has affected home values in California in recent years. While wildfire-prone areas offer scenery and green spaces that homeowners seek, measures of changing home values in recent years indicate that the risks may outweigh the benefits, even accounting for potential protection from homeowner’s insurance. This pattern may become stronger in years to come if residential construction continues to expand into areas with higher fire risk and if trends in wildfire severity continue.

Aylward, James, and Luiz E. Oliveira. 2020. “ Rising Wildfire Risk for the 12th District Economy .” FRBSF Economic Letter 2020-19 (July 13).

Donovan, Geoffrey H., Patricia A. Champ, and David T. Butry. 2007. “Wildfire Risk and Housing Prices: A Case Study from Colorado Springs.” Land Economics 83(2), pp. 217–233.

Mockrin, Miranda, Barbara McGuinness, David Helmers, and Volker Radeloff. 2023. Understanding the Wildland-Urban Interface (1990–2020) . Madison, WI: U.S. Department of Agriculture, Forest Service, Northern Research Station.

Mueller, Julie, John Loomis, and Armando González-Cabán. 2009. “Do Repeated Wildfires Change Homebuyers’ Demand for Homes in High-Risk Areas? A Hedonic Analysis of the Short and Long-Term Effects of Repeated Wildfires on House Prices in Southern California.” Journal of Real Estate Finance and Economics 38, pp.155–172.

Opinions expressed in FRBSF Economic Letter do not necessarily reflect the views of the management of the Federal Reserve Bank of San Francisco or of the Board of Governors of the Federal Reserve System. This publication is edited by Anita Todd and Karen Barnes. Permission to reprint portions of articles or whole articles must be obtained in writing. Please send editorial comments and requests for reprint permission to [email protected]

Democratic National Convention (DNC) in Chicago

Samantha Putterman, PolitiFact Samantha Putterman, PolitiFact

Leave your feedback

  • Copy URL https://www.pbs.org/newshour/politics/fact-checking-warnings-from-democrats-about-project-2025-and-donald-trump

Fact-checking warnings from Democrats about Project 2025 and Donald Trump

This fact check originally appeared on PolitiFact .

Project 2025 has a starring role in this week’s Democratic National Convention.

And it was front and center on Night 1.

WATCH: Hauling large copy of Project 2025, Michigan state Sen. McMorrow speaks at 2024 DNC

“This is Project 2025,” Michigan state Sen. Mallory McMorrow, D-Royal Oak, said as she laid a hardbound copy of the 900-page document on the lectern. “Over the next four nights, you are going to hear a lot about what is in this 900-page document. Why? Because this is the Republican blueprint for a second Trump term.”

Vice President Kamala Harris, the Democratic presidential nominee, has warned Americans about “Trump’s Project 2025” agenda — even though former President Donald Trump doesn’t claim the conservative presidential transition document.

“Donald Trump wants to take our country backward,” Harris said July 23 in Milwaukee. “He and his extreme Project 2025 agenda will weaken the middle class. Like, we know we got to take this seriously, and can you believe they put that thing in writing?”

Minnesota Gov. Tim Walz, Harris’ running mate, has joined in on the talking point.

“Don’t believe (Trump) when he’s playing dumb about this Project 2025. He knows exactly what it’ll do,” Walz said Aug. 9 in Glendale, Arizona.

Trump’s campaign has worked to build distance from the project, which the Heritage Foundation, a conservative think tank, led with contributions from dozens of conservative groups.

Much of the plan calls for extensive executive-branch overhauls and draws on both long-standing conservative principles, such as tax cuts, and more recent culture war issues. It lays out recommendations for disbanding the Commerce and Education departments, eliminating certain climate protections and consolidating more power to the president.

Project 2025 offers a sweeping vision for a Republican-led executive branch, and some of its policies mirror Trump’s 2024 agenda, But Harris and her presidential campaign have at times gone too far in describing what the project calls for and how closely the plans overlap with Trump’s campaign.

PolitiFact researched Harris’ warnings about how the plan would affect reproductive rights, federal entitlement programs and education, just as we did for President Joe Biden’s Project 2025 rhetoric. Here’s what the project does and doesn’t call for, and how it squares with Trump’s positions.

Are Trump and Project 2025 connected?

To distance himself from Project 2025 amid the Democratic attacks, Trump wrote on Truth Social that he “knows nothing” about it and has “no idea” who is in charge of it. (CNN identified at least 140 former advisers from the Trump administration who have been involved.)

The Heritage Foundation sought contributions from more than 100 conservative organizations for its policy vision for the next Republican presidency, which was published in 2023.

Project 2025 is now winding down some of its policy operations, and director Paul Dans, a former Trump administration official, is stepping down, The Washington Post reported July 30. Trump campaign managers Susie Wiles and Chris LaCivita denounced the document.

WATCH: A look at the Project 2025 plan to reshape government and Trump’s links to its authors

However, Project 2025 contributors include a number of high-ranking officials from Trump’s first administration, including former White House adviser Peter Navarro and former Housing and Urban Development Secretary Ben Carson.

A recently released recording of Russell Vought, a Project 2025 author and the former director of Trump’s Office of Management and Budget, showed Vought saying Trump’s “very supportive of what we do.” He said Trump was only distancing himself because Democrats were making a bogeyman out of the document.

Project 2025 wouldn’t ban abortion outright, but would curtail access

The Harris campaign shared a graphic on X that claimed “Trump’s Project 2025 plan for workers” would “go after birth control and ban abortion nationwide.”

The plan doesn’t call to ban abortion nationwide, though its recommendations could curtail some contraceptives and limit abortion access.

What’s known about Trump’s abortion agenda neither lines up with Harris’ description nor Project 2025’s wish list.

Project 2025 says the Department of Health and Human Services Department should “return to being known as the Department of Life by explicitly rejecting the notion that abortion is health care.”

It recommends that the Food and Drug Administration reverse its 2000 approval of mifepristone, the first pill taken in a two-drug regimen for a medication abortion. Medication is the most common form of abortion in the U.S. — accounting for around 63 percent in 2023.

If mifepristone were to remain approved, Project 2025 recommends new rules, such as cutting its use from 10 weeks into pregnancy to seven. It would have to be provided to patients in person — part of the group’s efforts to limit access to the drug by mail. In June, the U.S. Supreme Court rejected a legal challenge to mifepristone’s FDA approval over procedural grounds.

WATCH: Trump’s plans for health care and reproductive rights if he returns to White House The manual also calls for the Justice Department to enforce the 1873 Comstock Act on mifepristone, which bans the mailing of “obscene” materials. Abortion access supporters fear that a strict interpretation of the law could go further to ban mailing the materials used in procedural abortions, such as surgical instruments and equipment.

The plan proposes withholding federal money from states that don’t report to the Centers for Disease Control and Prevention how many abortions take place within their borders. The plan also would prohibit abortion providers, such as Planned Parenthood, from receiving Medicaid funds. It also calls for the Department of Health and Human Services to ensure that the training of medical professionals, including doctors and nurses, omits abortion training.

The document says some forms of emergency contraception — particularly Ella, a pill that can be taken within five days of unprotected sex to prevent pregnancy — should be excluded from no-cost coverage. The Affordable Care Act requires most private health insurers to cover recommended preventive services, which involves a range of birth control methods, including emergency contraception.

Trump has recently said states should decide abortion regulations and that he wouldn’t block access to contraceptives. Trump said during his June 27 debate with Biden that he wouldn’t ban mifepristone after the Supreme Court “approved” it. But the court rejected the lawsuit based on standing, not the case’s merits. He has not weighed in on the Comstock Act or said whether he supports it being used to block abortion medication, or other kinds of abortions.

Project 2025 doesn’t call for cutting Social Security, but proposes some changes to Medicare

“When you read (Project 2025),” Harris told a crowd July 23 in Wisconsin, “you will see, Donald Trump intends to cut Social Security and Medicare.”

The Project 2025 document does not call for Social Security cuts. None of its 10 references to Social Security addresses plans for cutting the program.

Harris also misleads about Trump’s Social Security views.

In his earlier campaigns and before he was a politician, Trump said about a half-dozen times that he’s open to major overhauls of Social Security, including cuts and privatization. More recently, in a March 2024 CNBC interview, Trump said of entitlement programs such as Social Security, “There’s a lot you can do in terms of entitlements, in terms of cutting.” However, he quickly walked that statement back, and his CNBC comment stands at odds with essentially everything else Trump has said during the 2024 presidential campaign.

Trump’s campaign website says that not “a single penny” should be cut from Social Security. We rated Harris’ claim that Trump intends to cut Social Security Mostly False.

Project 2025 does propose changes to Medicare, including making Medicare Advantage, the private insurance offering in Medicare, the “default” enrollment option. Unlike Original Medicare, Medicare Advantage plans have provider networks and can also require prior authorization, meaning that the plan can approve or deny certain services. Original Medicare plans don’t have prior authorization requirements.

The manual also calls for repealing health policies enacted under Biden, such as the Inflation Reduction Act. The law enabled Medicare to negotiate with drugmakers for the first time in history, and recently resulted in an agreement with drug companies to lower the prices of 10 expensive prescriptions for Medicare enrollees.

Trump, however, has said repeatedly during the 2024 presidential campaign that he will not cut Medicare.

Project 2025 would eliminate the Education Department, which Trump supports

The Harris campaign said Project 2025 would “eliminate the U.S. Department of Education” — and that’s accurate. Project 2025 says federal education policy “should be limited and, ultimately, the federal Department of Education should be eliminated.” The plan scales back the federal government’s role in education policy and devolves the functions that remain to other agencies.

Aside from eliminating the department, the project also proposes scrapping the Biden administration’s Title IX revision, which prohibits discrimination based on sexual orientation and gender identity. It also would let states opt out of federal education programs and calls for passing a federal parents’ bill of rights similar to ones passed in some Republican-led state legislatures.

Republicans, including Trump, have pledged to close the department, which gained its status in 1979 within Democratic President Jimmy Carter’s presidential Cabinet.

In one of his Agenda 47 policy videos, Trump promised to close the department and “to send all education work and needs back to the states.” Eliminating the department would have to go through Congress.

What Project 2025, Trump would do on overtime pay

In the graphic, the Harris campaign says Project 2025 allows “employers to stop paying workers for overtime work.”

The plan doesn’t call for banning overtime wages. It recommends changes to some Occupational Safety and Health Administration, or OSHA, regulations and to overtime rules. Some changes, if enacted, could result in some people losing overtime protections, experts told us.

The document proposes that the Labor Department maintain an overtime threshold “that does not punish businesses in lower-cost regions (e.g., the southeast United States).” This threshold is the amount of money executive, administrative or professional employees need to make for an employer to exempt them from overtime pay under the Fair Labor Standards Act.

In 2019, the Trump’s administration finalized a rule that expanded overtime pay eligibility to most salaried workers earning less than about $35,568, which it said made about 1.3 million more workers eligible for overtime pay. The Trump-era threshold is high enough to cover most line workers in lower-cost regions, Project 2025 said.

The Biden administration raised that threshold to $43,888 beginning July 1, and that will rise to $58,656 on Jan. 1, 2025. That would grant overtime eligibility to about 4 million workers, the Labor Department said.

It’s unclear how many workers Project 2025’s proposal to return to the Trump-era overtime threshold in some parts of the country would affect, but experts said some would presumably lose the right to overtime wages.

Other overtime proposals in Project 2025’s plan include allowing some workers to choose to accumulate paid time off instead of overtime pay, or to work more hours in one week and fewer in the next, rather than receive overtime.

Trump’s past with overtime pay is complicated. In 2016, the Obama administration said it would raise the overtime to salaried workers earning less than $47,476 a year, about double the exemption level set in 2004 of $23,660 a year.

But when a judge blocked the Obama rule, the Trump administration didn’t challenge the court ruling. Instead it set its own overtime threshold, which raised the amount, but by less than Obama.

Support Provided By: Learn more

Educate your inbox

Subscribe to Here’s the Deal, our politics newsletter for analysis you won’t find anywhere else.

Thank you. Please check your inbox to confirm.

research paper on fire ecology

  • Open access
  • Published: 14 June 2023

The scientific value of fire in wilderness

  • Mark R. Kreider   ORCID: orcid.org/0000-0002-1518-8267 1 ,
  • Melissa R. Jaffe 1 ,
  • Julia K. Berkey 2 ,
  • Sean A. Parks 3 &
  • Andrew J. Larson 1  

Fire Ecology volume  19 , Article number:  36 ( 2023 ) Cite this article

3170 Accesses

5 Citations

8 Altmetric

Metrics details

Wilderness areas are important natural laboratories for scientists and managers working to understand fire. In the last half-century, shifts in the culture and policy of land management agencies have facilitated the management practice of letting some naturally ignited fires burn, allowing fire to fulfill its ecological role and increasing the extent of fire-related research opportunities. With the goal of identifying the global scientific advances enabled by this paradigm shift in wilderness fire management, we conducted a systematic review of publications that either (1) selected protected areas for investigation because of an active fire regime enabled by wilderness fire management, (2) studied modern fires or fire regimes deliberately located in a wilderness area, or (3) conducted applied research to support wilderness fire management.

Our systematic review returned a sample of 222 publications that met these criteria, with an increase in wilderness fire science over time. Studies largely occurred in the USA and were concentrated in a relatively small number of protected areas, particularly in the Northern Rocky Mountains. As a result, this sample of wilderness fire science is highly skewed toward areas of temperate mixed-conifer forests and historical mixed-severity fire regimes. Common principal subjects of publications included fire effects (44%), wilderness fire management (18%), or fire regimes (17%), and studies tended to focus on vegetation, disturbance, or wilderness management as response variables.

Conclusions

This work identifies major scientific contributions facilitated by fire in wilderness, including self-limitation of fire, the effects of active fire regimes on forest and aquatic systems, barriers and potential solutions to wilderness fire management, and the effect of fire on wilderness recreation and visitor experiences. Our work reveals geographic and bioclimatic areas where more research attention is needed and highlights under-represented wilderness areas that could serve to fill these gaps. Finally, we identify priorities for future wilderness fire research, including the past and potential role of Indigenous and prescribed burning, the effects of changing climate and fire regimes on ecosystem processes, and how to overcome barriers to wilderness fire management.

Antecedentes

Las áreas silvestres son laboratorios naturales importantes para científicos y gestores que trabajan para entender el fuego. En los pasados 50 años, cambios en la cultura y en políticas de manejo del fuego promovidas por las agencias de manejo de tierras, han facilitado la práctica de dejar que algunos fuegos iniciados naturalmente (i.e. por rayos) puedan quemar diferentes superficies, permitiendo que el fuego cumpla con su rol ecológico e incrementar asimismo las oportunidades de extender las investigaciones relacionadas con el fuego. Con el objetivo de identificar los avances científicos globales alcanzados por este cambio de paradigma en el manejo del fuego en áreas silvestres, condujimos una revisión sistemática de publicaciones que: (1) eligieron áreas protegidas para la investigación pues tenían un régimen activo de fuegos permitido por el manejo del fuego en esas áreas silvestres, (2) estudiaban fuegos modernos o actuales o regímenes de fuego ubicados deliberadamente en un área silvestre, o (3) condujeran investigación aplicada para apoyar el manejo del fuego en esas áreas silvestres.

Nuestra investigación sistemática condujo a una muestra de 222 publicaciones que cumplía con esos criterios, con un incremento paulatino, en el tiempo, de investigaciones sobre la ciencia del fuego en áreas silvestres. Los estudios fueron en su mayoría originados en los EEUU y estuvieron concentrados en un número reducido áreas protegidas, y particularmente en las montañas rocallosas del norte. Como resultado, esta muestra de la ciencia del fuego en áreas silvestres está totalmente sesgada hacia bosques templados mixtos de coníferas que tenían históricamente regímenes de fuego de severidad mixta. Los sujetos principales de publicación incluían efectos del fuego (44%), manejo del fuego en áreas silvestres (18%) o regímenes de fuego (17%), y los estudios tendían a enfocarse en la vegetación, los disturbios, o en el manejo de áreas silvestres como variable de respuesta.

Conclusiones

Este trabajo identifica contribuciones científicas importantes facilitadas por los incendios en áreas silvestres, incluyendo la autolimitación del fuego, los efectos de los regímenes activos del fuego en el bosque o en los sistemas acuáticos, las barreras y soluciones potenciales al manejo del fuego en áreas silvestres, y los efectos del fuego en la recreación y experiencias de los visitantes de estas áreas. Nuestro trabajo revela áreas geográficas y bioclimáticas donde mayor atención debe ser puesta en relación a la necesidad de realizar investigaciones en fuegos, y destaca áreas silvestres que podrían servir para llenar esos vacíos. Finalmente, identificamos prioridades para futuras Investigaciones en fuego en áreas silvestres, incluyendo el rol pasado y potencial de los indígenas y de las quemas prescriptas, los efectos del cambio climático y regímenes de fuego en procesos ecosistémicos, y como superar barreras en el manejo del fuego en áreas silvestres.

Wilderness and other protected areas provide value to society as places for scientific research and knowledge production. This is true in a strict sense for congressionally designated Wilderness Areas in the United States of America (USA), where the Wilderness Act of 1964 explicitly identifies scientific use as one of the six public purposes of wilderness (“Wilderness Act 16 U.S. Code § 1131,” 1964 ). More generally, scientific study of ecosystems in wilderness and protected areas provides the basis for developing natural models of ecosystem structure and dynamics, including the role of natural disturbances (Franklin et al. 2002 ; Berkey et al. 2021a ). This knowledge informs ecosystem restoration and conservation (Hopkins et al. 2014 ), including the development of ecologically based management systems used outside of formal reserves (Kuuluvainen et al. 2021 ).

A profound contribution of wilderness and protected area management has been to catalyze a paradigm shift from fire suppression to fire management for resource benefit (Van Wagtendonk 2007 ). This is especially true in the USA, where, for much of the 20 th century, there was very little fire activity due to the 10 AM Policy—a national policy enacted in 1935 to suppress all wildfire ignitions—as well as earlier depopulation and displacement of Native Americans and their use of fire (Fisher 1997 ; Kimmerer and Lake 2001 ; Ostlund et al. 2005 ; Roos et al. 2021 ). However, the Leopold report (Leopold et al. 1963 ), which stimulated the National Park Service to recognize fire as an ecological process (Rothman 2007 ), along with the Wilderness Act (“Wilderness Act 16 U.S. Code § 1131,” 1964 ), which prompted Forest Service managers in the US Northern Rocky Mountains to manage some natural ignitions (Smith 2014 ; Berkey et al. 2021b ), began to restore fire as an ecological process and management tool in some parks and wilderness areas starting in the late 1960s and early 1970s. At this same time, a shift was also occurring in the scientific literature, acknowledging the important role of fire in ecosystems (Habeck and Mutch 1973 ; Heinselman 1973 ; Kilgore 1973 ; Wright 1974 ).Together, these changes created opportunities to study fire as an socioecological process and required development of new knowledge to support fire management decision making (Agee 2000 ; Kilgore 1987 ; Miller and Aplet 2016 ; Smith 2014 ).

We assessed the scientific contributions and knowledge production enabled by the shift toward recognizing fire as an integral ecosystem process, and the accompanying development of wilderness fire management in some places. Our review is partially motivated by the Wilderness Act’s explicit identification of scientific use as one of the purposes of wilderness. Wilderness has, in the past, been criticized as not delivering on the promise and potential as a place for research (Franklin 1987 ); we question if that holds in the case of wilderness fire science in the present. While wilderness is largely a legal and philosophical construct originating from the early and middle 20 th century environmental protection movement in the USA, many protected areas globally have active fire regimes. To include scientific contributions from those regions, we defined the geographic scope of our study to be global. Our specific objectives were to:

Summarize the scientific contributions made possible by wilderness fires and wilderness fire management in terms of their distribution in time and space, principal subject and environmental resource, and type of study and publication.

Assess the representativeness of studies in our sample in climate and fire regime space.

Synthesize major areas of scientific advancement and discovery made possible by wilderness fire management and identify future research priorities.

To establish the scope of our review, we defined wilderness as protected areas globally where natural disturbance processes such as fire are allowed to proceed under some cases. We thus used International Union for Conservation of Nature (IUCN) protected area management categories Ia (strict nature reserve), Ib (wilderness), and II (national park) (Dudley 2013 ). Though most naturally-ignited fires in wilderness are suppressed to some extent (Miller 2012 ), these areas nonetheless tend to have less suppression than outside of wilderness (Haire et al. 2013 ; Morgan et al. 2014 ), and are not subject to intensive management such as salvage logging.

Database search : We conducted a database search to identify a global sample of studies where fire in wilderness created either the opportunity or the need for research. Initially, we tested several search strings, including [“Wilderness” AND “fire”], [“National Park” AND “fire”], [“National Wildlife Refuge” AND “fire”], [“National Preserve” AND “fire”], and [“National Monument” AND “fire”], as well as searches for individual wilderness areas, national parks, or regions [e.g., “Denali National Park” AND “fire”]. Preliminary analysis of these search strings revealed that searches other than [“Wilderness” AND “fire”] were overly sensitive, returning many studies that did not meet our inclusion criteria. Thus, we ultimately compiled our dataset from a sample of the literature using the single search string [“Wilderness” AND “fire”]. These preliminary and final searches took place during May 2019 using the ISI Web of Science ( https://webofknowledge.com ) and  U.S. Forest Service Treesearch ( https://www.fs.usda.gov/treesearch/ ) databases.

We screened all publications, retaining those that met at least one of the following criteria: 1) studies that had selected a wilderness or other protected area for investigation because of the modern (post-mid-20 th century) active fire regime enabled by wilderness fire management; 2) studies of modern fires or fire regimes deliberately located in wilderness or other protected areas; 3) applied research undertaken to support implementation or continuation of wilderness fire management. We used systematic literature review methods (Pullin and Stewart 2006 ) and placed no disciplinary or subject matter constraints on our review—our objective was to document the full range of scientific contributions made possible by wilderness fire management. However, we did exclude studies conducted in wilderness but with a pre-historical or historical focus prior to the mid-20 th century. We also excluded large scale (e.g., regional to subcontinental scale) studies where the inclusion of protected areas was incidental to the core focus or study area. We retained reviews, syntheses, and meta-analyses when the scope, inference, or conclusions of these publications depended significantly on the contribution of one or more qualifying (as described above) wilderness fire studies. Four of the authors (MRK, MRJ, SAP, AJL) assessed publications for inclusion. We automatically included publications when three or more reviewers independently recommended inclusion in the final dataset, with ties reassessed and decided by the senior author.

To identify the scientific advances within this final dataset, we collected information on each study’s research subject, themes, and location. The same four authors each assessed every publication in the final dataset to collect information on publication type, study type, principal subject, environmental resource, country, and protected area (Appendix 1 ). We initially used the Joint Fire Science Program (JFSP) Findings Data Dictionary ( https://www.firescience.gov/PSR/documents/Findings_Data_Dictionary.pdf ) to define possible categories for the study type, principal subject, and environmental resource attributes. However, preliminary review of qualifying studies in our sample showed a greater breadth of environmental resources and study types than those listed in the JFSP data dictionary. Thus, we ultimately adopted the value definitions described in Appendix 2 for definitions of possible publication type, study type, principal subject, and environmental resource categories.

Representativeness: To assess how representative our sample was of broader climate and fire activity, we compared patterns of climate and historical fire regimes represented in sampled wilderness areas to those of 1) wilderness areas in general and 2) all land designations. Because most studies focused on protected areas in the contiguous United States, we restricted our representativeness analyses to this area.

To assess the climatic representativeness of sampled areas, we constructed climate envelopes using annual climate water deficit and actual evapotranspiration data (aggregated to 1981–2010 averages) from gridded TerraClimate datasets (Abatzoglou et al. 2018 ). We compared the climate envelope for sampled wilderness areas to 1) a climate envelope of all wilderness areas in the contiguous USA and to 2) a climate envelope of the entire contiguous USA. To assess the historical fire regime representativeness of sampled areas, we constructed fire regime envelopes using Mean Fire Return Interval (MFRI) and Percent of Replacement-Severity Fire (PRS) from gridded LANDFIRE datasets (Rollins 2009 ). We converted these binned categorical values to their average value (e.g., the Replacement-Severity Fire category of 41-45% was converted to 43%). As before, we compared the fire regime envelope for sampled areas to that of all wilderness areas in the contiguous USA, as well as to that of the entire contiguous USA. We accessed TerraClimate and LANDFIRE datasets via Google Earth Engine (Gorelick et al. 2017 ), and extracted the values of all pixels at 4-km scale that fell within sampled wilderness areas, contiguous USA wilderness areas, and the entire contiguous USA respectively.

Our initial keyword search returned 608 publications. Following the screening process, 222 publications were retained in our final sample and analyzed (Appendix 3 ). Code and data to reproduce all results and figures from this paper can be accessed through the Zenodo open-access repository at https://doi.org/10.5281/zenodo.6326355 .

Summary statistics

Most studies in our sample reported on research conducted in the USA (90%). Australia (6%) and Canada (5%) were the only other countries with more than one publication, with a handful of additional countries—Dominican Republic, Mongolia, Russian Federation, South Africa, Spain, Zambia, and Zimbabwe—each the subject of a single publication (Fig. 1 ). Percentages sum to greater than 100 because nine publications focused on more than one country. Publications in our sample were published from 1970–2019, with an increasing trend in publications per year through time (Fig. 2 ).

figure 1

A  Number of studies taking place in each country. Note that some studies ( n = 9) reported on research in more than one country. B  Frequency of studies by wilderness area (USA only). Of the 199 studies from the USA in our sample, none documented research outside of the contiguous USA. Labels shown for the 10 wilderness areas with the most studies (Bob Marshall Wilderness, Scapegoat Wilderness, and Great Bear Wilderness were combined into “Bob Marshall Wilderness Complex”; Gila Wilderness and Aldo Leopold Wilderness were combined into “Gila / Aldo Leopold Wilderness Complex”). Note that many studies occurred in multiple wilderness areas

figure 2

A  Frequency of studies by year. B  Cumulative frequency of the 10 wilderness areas with the most studies. Circles indicate the first year the wilderness area occurs in our sample. Bob Marshall Wilderness, Scapegoat Wilderness, and Great Bear Wilderness were combined into “Bob Marshall Wilderness Complex”; Gila Wilderness and Aldo Leopold Wilderness were combined into “Gila / Aldo Leopold Wilderness Complex”

Most publications in our final sample were journal articles (68%), with proceedings papers another common avenue for wilderness fire science (18%). The remaining publications were from books or book chapters (5%), General Technical Reports (3%), datasets (2%), management documents (2%), or other (2%) (Fig. 3 A). Publications spanned many study types (Fig. 3 B), with most reporting on new data in the form of observational studies (62%) or synthesizing information through reviews/meta-analyses (25%). The remainder of publications were modeling studies (7%), methods papers (3%), datasets (2%), or field experiments (1%).

figure 3

Percentage of studies by A  publication type, B  study type, C  principal subject, and D  environmental resource. *Because studies could have more than one Environmental resource, values sum to greater than 100%

Publications in our sample focused on a variety of principal subjects (Fig. 3 C). The most common were publications primarily dealing with fire effects (44%), with additional representation from incident management (18%), fire regimes (17%), and fire ecology (12%). Remaining publications focused on fuel treatments (5%), monitoring (2%), fire behavior (1%), tool assessment (<1%), smoke management (<1%), and fuel characterization (<1%). Beyond their primary focus, publications dealt with an even more varied suite of environmental resources, or response variables. Over half of publications explored fire effects on vegetation (64%), patterns of fire (57%), or wilderness management in the context of active fire management (51%). Publications also reported, in lower numbers, on a wide variety of other response variables (Fig. 3 D). Because publications could have more than one response variable, percentages sum to more than 100.

Publications in our sample that focused on fire ecology and fire effects were more likely to be published in peer-reviewed journals, while publications that focused on fire regimes, incident management, and fuel treatments were more likely to be published in proceedings papers (Fig. 4 ). Principal subjects of publications also tended to be linked to specific types of environmental resources. For example, fire ecology, fire effects, and fire regime publications focused more often on physical variables such as soil, water, vegetation, and biota, while publications with principal subjects of fuel treatment or incident management focused on more abstract variables such as economics or law/policy (Fig. 5 ).

figure 4

Proportion of publication type by principal subject. Only the five principal subjects with the most publications are shown ( n = 212; 95% of studies). Numbers on top of each column indicate the number of studies in that category

figure 5

Connections between the principal subjects and environmental resources of publications

Representativeness

All the publications from the United States of America ( n = 199) occurred in the contiguous USA (Fig. 1 ), and we conducted further analysis of representativeness on this sub-sample. Within the USA, studies were largely concentrated in the Northern Rocky Mountains, several southwestern wilderness areas, the Sierra Nevada, and the Boundary Waters Canoe Area Wilderness (Fig. 1 ). Climate of wilderness areas represented in our sub-sample occupied a reduced climate envelope (Fig. 6 E) compared both to wilderness areas in the contiguous USA (Fig. 6 C) and especially the contiguous USA at large (Fig. 6 A). Research from this sub-sample has predominately occurred in areas with climate characterizing mixed-conifer forests.

figure 6

Climate and fire regime envelopes for the contiguous USA A , B ; all wilderness areas in the contiguous USA C , D ; and only wilderness areas in our sample E , F . Grey shading in the lefthand maps show the spatial extent of pixels contributing to each row. Envelopes are approximated by 2D density plots (orange) with actual values shown by black dots. Data in E and F are proportional to the number of times a wilderness area was included in the sample (i.e., if a wilderness area was included 10 times in the sample, each pixel value from that wilderness area is also included 10 times)

In a similar manner, historical fire regimes of studied wilderness areas (Fig. 6 F) represent a reduced fire regime envelope relative to wilderness areas in the contiguous USA (Fig. 6 D) and the contiguous USA overall (Fig. 6 B). Historical fire regimes of studied wilderness areas were clustered in mixed-severity regime space (i.e., stand-replacing proportion ~0.5 and mean return intervals of 30–100 years). There were few studied wilderness areas with historical frequent low-severity fire regimes, and virtually none with frequent stand-replacing fire regimes (i.e., grassland and shrubland ecosystems).

In the contiguous USA, every wilderness area with extensive fire in the last several decades (i.e., cumulative area >200,000 ha burned 1984–2019) is represented by at least one study in our sample (Fig. 7 A). However, many of the wilderness areas with little or no representation in our sample have, in fact, experienced a relatively high amount of fire since 1984 (Fig. 7 B).

figure 7

A  Relationship between total amount of fire burned (1984–2019) in each wilderness area in the contiguous USA and the number of times that wilderness area was studied in our sample. B  Inset of wilderness areas falling within the red box in panel A

Beyond the quantifiable metrics of research described above, we identified major conceptual areas in which scientific advancements have been facilitated by fire in wilderness. We do not imply that research from outside of wilderness areas has nothing to offer, but rather that the following advancements have depended, in significant part, on research opportunities afforded by wilderness fires and wilderness fire management. We also propose high-priority research questions which future wilderness fire science is well-suited to address.

Self-limitation

A primary scientific advancement enabled by wilderness fire management is the extent to which fire limits the spread and intensity of subsequent fire. Ecological theory of this pattern-process relationship between fire and vegetation (Agee 1999 ; Peterson 2002 ; Turner 1989 ) has been demonstrated with field data largely arising from studies in wilderness areas (e.g., Collins et al. 2009 ; Parks et al. 2016 , 2015 , 2014 ; Teske et al. 2012 ). Areas with a management history of wildland fire use are essential for this research (Miller and Aplet 2016 ), because locations with heavy suppression provide few instances of interactions between fire perimeters through time. Wilderness fire science has also revealed that the self-limiting effects of fire vary by ecosystem, diminish over time, and are reduced by extreme fire weather (Collins et al. 2009 ; Parks et al. 2015 ). This body of research underscores how the decision to suppress a fire is a lost opportunity to create natural fuel breaks and restore ecosystem resilience (Miller 2012 ; Parks et al. 2015 ). Wilderness areas with active fire regimes can serve as excellent places for future research that tests how changing climate and fire regimes will impact the strength and longevity of self-limitation effects following fire.

Forest ecosystem dynamics under active fire regimes

With high levels of fire suppression in nearly all non-wilderness areas (Calkin et al. 2005 ; Quadrennial Fire Review 2014 ), wilderness areas with active fire management offer some of the only contemporary insights into how active fire regimes (i.e., where fires are allowed to burn under a wider range of conditions) shape forest ecosystems. Research in wilderness areas has highlighted fire as a driver of heterogeneity, both by increasing structural complexity in forest ecosystems (e.g., Holden et al. 2006 ; Kane et al. 2013 ; Robinson et al. 2005 ) as well as by catalyzing shifts in composition that maintain dynamic landscape mosaics (e.g., Jackson and Sullivan 2009 ; Kleindl et al. 2015 ; Reilly et al. 2006 ; van Wagtendonk et al. 2012 ). Additionally, wilderness fire research has assessed the ability of wildfires to restore target ranges of structure and composition, showing that wildfires, especially when allowed to burn under less extreme conditions, can be a successful restoration treatment in ecosystems with low- and moderate-severity fire regimes (Fulé and Laughlin 2007 ; Larson et al. 2013 ; Pawlikowski et al. 2019 ; Taylor 2010 ). As fire activity increases in many areas (e.g., Schoennagel et al. 2017 ; Jain et al. 2022 ), research from wilderness areas provides an important ecological baseline (Belote et al. 2015 ; Frelich 2017 ), helping us to create mechanistic predictions of how ecosystems may respond to changing climate and fire regimes. Furthermore, wilderness areas provide an excellent opportunity to test whether locations with active fire regimes—which tend to have reduced fuels, more structurally-diverse forests, and greater landscape heterogeneity—exhibit greater resiliency or smoother transitions to ecological change than areas where fire continues to be suppressed (Coop et al. 2020 ).

Aquatic ecosystem dynamics under active fire regimes

Though representing a much smaller proportion of our sample relative to publications dealing with vegetation, an important body of wilderness fire science has advanced understandings of the effects of fire on fluvial geomorphology and aquatic processes and biota. Wildfires strongly influence the routing of wood and sediment from upland and riparian areas to the channel network (Robinson et al. 2005 ; Marcus et al. 2011 ; Kleindl et al. 2015 ), which can increase spatial complexity (Arkle et al. 2010 ; Robinson et al. 2005 ), shift species composition of macroinvertebrates (Arkle et al. 2010 ; Jackson et al. 2012 ; Jackson and Sullivan 2009 ; Malison and Baxter 2010 ), and provide salmonid spawning habitat (Jacobs et al. 2021 ). Physical changes associated with increased flows following wildfires may also have negative effects such as increased nutrient loadings (Spencer et al. 2003 ) and decreased abundance of macroinvertebrates and fish (Bozek and Young 1994 ; Minshall et al. 2001 ; Rugenski and Minshall 2014 ). However, restoring natural wildfire regimes can provide numerous benefits, including increasing snowpack and reducing forest water stress (Boisramé et al. 2019 ). Wilderness fire management provides many research opportunities to explore how changing climate and fire regimes will impact aquatic systems, and whether aquatic systems within an active fire regime are better able to adapt to these changes.

Wilderness fire management decision making

Advancements in our understanding of self-limitation have equipped wilderness managers with improved tools for predicting when wildfires can be safely managed within wilderness boundaries (e.g., Barnett et al. 2016 ; Scott et al. 2012 ; Suffling et al. 2008 ). A sizeable body of publications in our sample have also identified the social and institutional challenges to restoring natural fire regimes to wilderness areas, such as a poor public perception of fire, negative smoke impacts, and a lack of institutional support (e.g., Miller 2003 ; Miller et al. 2011 ; Parsons 2000 ; Parsons et al. 2003 ; Williamson 2007 ). As a result of these barriers, the majority of fires continue to be suppressed in all but a handful of wilderness areas, where historical precedents exist for allowing wilderness fire (Seielstad 2015 ; Berkey et al. 2021b ). To incentivize the wider implementation of active fire management, it is vital to increase public understanding of the inevitability of fire events and the importance of fire to ecosystem processes, build cooperation across administrative boundaries, and create a culture within land management agencies that equips, supports, and expects managers to manage fires for resource benefit when possible (Berkey et al. 2021b ; Miller et al. 2011 ). Fifty years of training and experience might be expected to have made it easier to manage wilderness fires for resource benefit, but instead, social and institutional barriers continue to discourage the practice on a widescale basis (Seielstad 2015 ). Future wilderness fire science  can investigate how to reverse this trend.

Fire impacts on recreation

Many of the qualities that draw recreationalists to wilderness areas—remoteness and ruggedness—are also what promote wildfires that burn with minimal, or no, suppression. As such, wilderness fire management often provides the opportunity for social science research exploring the effects of fire on recreation and recreationists’ experiences and attitudes. Wilderness fire science has shown that fires and resulting trail closures can negatively impact experiences of wilderness recreationalists (Boxall et al. 1996 ; Brown et al. 2008 ; Tanner et al.  2022 ), although not all wildfires have this adverse effect (e.g., Love and Watson 1992 ). In fact, visual evidence of disturbances can increase wilderness character and public interest (Schroeder and Schneider 2010 ), leading recreationalists to seek out recently burned areas (Englin et al. 2008 ; Dvorak and Small 2011 ; Sánchez et al. 2016 ; but see Tanner et al. 2022 ). Studies in wilderness have shown that recreationalists often support natural and prescribed fires in wilderness (Borrie et al. 2006 ; Knotek et al. 2008 ; McCool and Stankey 1986 ; Watson et al. 2015 ), although this support may drop during high fire-activity years (Borrie et al. 2006 ). Public support is crucial to the feasibility of wilderness fire management, and an important future role of wilderness fire science will be to understand how the increasing size and severity of wildfires (Schoennagel et al. 2017 ) is impacting patterns of recreation and recreationalists’ perceptions of wilderness fire management.

This systematic review illustrates how fire in wilderness has created opportunities for research—and therefore the production of knowledge—related to patterns, processes, and effects of wildfire, as well as management of wildfire. While we present a diversity of research topics and advancements that have originated from wilderness fire science, our analysis also reveals areas—geographic, bioclimatic, and conceptual—where more research attention is needed.

Our sample of wilderness fire science is heavily skewed towards studies from the contiguous USA. Less than 10% of studies in our sample reported on findings from outside of North America, even though many other regions of the world have experienced more fire over the last several decades (Robinne et al. 2019 ). We only searched for publications in English, and our search string of [“Wilderness” AND “Fire”] may have contributed to the observed bias by identifying fewer studies from countries with protected areas named with other descriptors (e.g., “Reserve”, “National Park”, “Provincial Park”, “Strictly Protected Area”, etc.). However, given that the USA had among the earliest adoption of wilderness fire management and provided an early model of wilderness areas as a construct, it is perhaps not surprising that many of these publications come from landscapes in the USA.

Within the USA, studies were also heavily concentrated in a relatively small number of wilderness areas, particularly in the northern Rockies. This pattern is largely driven by where wilderness managers have allowed fire to burn (Miller and Aplet 2016 ). For example, the Selway-Bitterroot Wilderness had the most studies in our sample by far, likely because it was the first US Forest Service-managed area to allow for scientific observation of fire (Smith 2014 ), as well as the first Forest Service area to adopt wilderness fire management (Berkey et al. 2021b ). Furthermore, fires are almost always suppressed in small wilderness areas (Zimmerman et al. 2006 ) because unplanned ignitions are more likely to spread outside of wilderness boundaries (Barnett et al. 2016 ). For this reason, large wild areas (e.g., Selway-Bitterroot Wilderness, Frank Church River of No Return Wilderness, Bob Marshall Wilderness Complex, Yellowstone National Park, and Gila/Aldo Leopold Wilderness Complex) have allowed for greater use of wilderness fire management, resulting in increased research attention.

Our study is a sample of a broader body of research, and thus does not capture every wilderness fire science study. For example, there are some wilderness areas which did not appear in our sample, but where research has occurred: e.g., Kalmiopsis Wilderness in Oregon (Thompson and Spies 2009 ; Donaghy Cannon 2013 ) and Ventana Wilderness in California (Talley and Griffin 1980 ). Furthermore, our search strings may have not detected studies in designated wilderness areas where the location is better known by another name (e.g., a study conducted in Marjory Stoneman Douglas Wilderness in Florida, but using “Everglades National Park” to describe the study location; Beckage et al. 2003 ; Ruiz et al. 2013 ). However, despite the imperfect detection of all wilderness fire science, we expect that the frequency with which wilderness areas appear in our sample is a useful proxy for the relative amounts of fire research attention, at least in the USA.

This sample of wilderness fire science is not fully representative of climate or fire regimes in the USA, and certainly not globally (Robinne et al. 2019 ). Rather, the sample is highly skewed toward the climate space of temperate mixed-conifer forests and the fire regime space of mixed-severity fire—largely due to where designated wilderness areas occur. Even if all current wilderness areas in the contiguous USA had active fire regimes, knowledge derived from these areas would still represent a reduced climate and fire regime space relative to the whole country (Fig. 6 ). Nevertheless, even when only considering available wilderness areas, there is potential to broaden the scope of fire science to better include under-represented climates and historical fire regimes. We identify many wilderness areas that have experienced significant wildfire but where our database search detected little or no research (e.g., many of the labeled wilderness areas in Fig. 7 B). These under-represented areas offer the possibility for studies that would expand the geographic, climate, and fire regime spaces of wilderness fire science, thereby helping to address knowledge gaps. Additionally, allowing more fire to burn in wilderness areas with little to no contemporary fire can create additional research opportunities, especially in wilderness areas that might help to expand the representativeness of the current body of wilderness fire research.

While the body of research documented here covers a diversity of research areas and questions, there are several notable conceptual gaps. First, wilderness fire science has largely focused on fire effects on vegetation. Notwithstanding the valuable advances in other categories described in the synthesis section above, we urge the continued increase of research that explore how fire impacts other domains such as wildlife, fungi, the pyrodiversity-biodiversity hypothesis, soil, aquatic systems, and human dimensions in a wilderness context. Secondly, very few publications in our sample explicitly addressed the impacts of climate change on fire dynamics in wilderness (e.g., Frelich and Reich 2009 ; Rugenski and Minshall 2014 ). Many of the high-priority future research areas identified in the synthesis section relate to climate change, and we urge the greater use of wilderness areas as a natural laboratory to explore impacts of climate change on fire regimes and fire-prone ecosystems (Belote et al. 2015 ). Finally, while there is a growing appreciation that most wilderness areas were historically managed and impacted by Indigenous groups and their use of fire (Fisher 1997 ; Kimmerer and Lake 2001 ; Watson et al. 2011 ), our sampled detected very few publications that focused on the past or present role of cultural or Indigenous burning in wilderness (e.g., Kay 2000 ; Trauernicht et al. 2013 ). Despite the fact that Indigenous burning was identified as a research focus already forty years ago at a large North American symposium on wilderness fire (Kilgore 1987 ; Lotan et al. 1985 ), we found that this emphasis has largely diminished in our sample in more recent years. Research published since our analysis has begun to address this gap (e.g., Kipfmueller et al. 2021 ; Larson et al. 2021 ), however we believe that wilderness areas provide an excellent opportunity to expand on this vital research area.

Another critical issue that remains unresolved is the role of prescribed fire in wilderness settings. Prescribed fire is legally allowed in virtually all wilderness areas, with varying goals of reducing fire hazard, restoring historical structure and habitat, or allowing natural fire regimes to return (e.g., Bureau of Land Management, 2012 ; National Park Service, 2006 ; U.S. Fish and Wildlife Service 2008 ; U.S. Forest Service 2007 ). Furthermore, many backcountry recreationalists support the use of prescribed fire in wilderness (Knotek et al. 2008 ; McCool and Stankey 1986 ). However, prescribed fire is rarely implemented and is subject to many of the same challenges inherent in managing wildfire (Jaffe et al. 2020 ; Parsons 2000 ; Schultz et al. 2018 ) in addition to philosophical questions as to the appropriate level of human influence in wilderness (Lawhon 2011 ; Lotan et al. 1985 ; Parsons et al. 2003 ). Several publications identified by our review have attempted to overcome these hurdles: e.g., by making an ecological case for prescribed fire (Heinselman 1970 ), showing the important role that prescribed fire can play in restoring degraded habitats (Keane et al. 2006 ; Vequist 2007 ), or demonstrating that prescribed fire can often meet many of the management objectives of wildfires (Nesmith et al. 2011 ). However, the limited application of prescribed fire in wilderness areas today shows that additional work is needed, including applied research to help understand perceived policy barriers to using prescribed fire in wilderness and develop strategies to overcome these barriers. Future wilderness fire science can explore the extent to which prescribed fire emulates natural patterns of wildfire (e.g., seasonality, severity, duration, patch size), especially under changing climate and fire regimes.

Unlike previous reviews of wilderness fire science (Agee 2000 ; Kilgore 1987 ; Miller and Aplet 2016 ), we used systematic methods, allowing for quantitative analysis of research patterns. Additionally, our study differs from previous reviews in that we maintained a global scope. The previous reviews explicitly acknowledge their general emphasis on fire science from the western USA (e.g., Kilgore 1987 ; Miller and Aplet 2016 ), and while our study clearly bears witness to and quantifies this bias, it also highlights important research from other parts of the USA and globally. In the nearly four decades spanning these reviews, wilderness fire science has helped to identify and address many important research questions (Table 1 ). For example, fire as a landscape process and as a driver of complexity have remained active topics of wilderness fire science for several decades, as knowledge of these areas is continually deepened and expanded. Improved data and models were key areas featured by both Agee ( 2000 ) and Miller and Aplet ( 2016 )—however, because these advances developed from and apply to fire science broadly, they were largely outside the defined scope of our wilderness-specific study and search strings.

Identifying and overcoming barriers to wilderness fire management has been a focus in every wilderness fire science review (Table 1 ), highlighting the complex and persistent hurdles faced by wilderness fire managers and calling into question whether this problem will be solely solved by additional research (Miller and Aplet 2016 ). Successful implementation of wilderness fire management depends both on managers with a deep commitment to fire as a fundamental ecological process, as well as strong support from institutional leaders to deal with the short-term risk incurred by allowing fire (Berkey et al. 2021b ). It is important to support initiatives that cultivate and deepen managers’ professional wilderness fire ethic and encourage leaders to provide a supportive environment where this ethic can be expressed instead of being at odds with institutional leadership.

Some topics featured in wilderness fire science reviews have experienced a resurgence in recent years (Table 1 ). The roles of Indigenous and prescribed fire in wilderness were a major focus of the field at the time of Kilgore’s review ( 1987 ), however they are absent from the priority research areas of later reviews (Agee 2000 ; Miller and Aplet 2016 ). Today, Indigenous and prescribed fire will likely play an important role in addressing the impacts of changing climate and fire regimes in protected areas, and increased research and discussion is vital (Larson et al. 2021 ).

Our review identified over 220 scientific studies enabled by fire in wilderness. Given that we were focused on the relatively narrow topic of fire, our sample of scientific literature is a conservative estimate of the total scientific contribution of wilderness. Research and scientific use of wilderness is often questioned and challenged by managers (Landres 2010 ), and the policies of some agencies force researchers to demonstrate that the work cannot be accomplished outside of wilderness. This distinction is not mandated by law or required of other wilderness uses or user groups (e.g., a recreational visitor does not need to demonstrate that their recreational activity can only be accomplished in wilderness before they are allowed to visit). Greater effort to quantify the scope, impact, and societal benefits of scientific research and monitoring conducted in wilderness, or in support of wilderness management, could help wilderness managers better understand the role of wilderness in larger socioecological systems (Parsons 2007 ; Smith and Gray 2021 ), potentially leading to greater support from managers for scientific activities in wilderness. At the same time, researchers have a responsibility to familiarize themselves with wilderness law and policy—especially section 4c of the Wilderness Act (“Wilderness Act 16 U.S. Code § 1131,” 1964 ) which describes prohibited uses—as well as important wilderness management decision support tools, such as minimum requirements analysis. This will help to ensure that researchers propose appropriate methods, reducing conflicts with wilderness managers.

Wilderness fire science has increased in pace and scope over the last five decades, helping to advance knowledge in a variety of conceptual areas, including self-limitation of fire, forest and aquatic ecosystem dynamics under active fire regimes, fire management and decision-making, and fire effects on recreation and visitor experiences. Systematic methods enabled us to detect a wide range of disciplines; however, we show that our sample of wilderness fire science was heavily skewed towards studies from a handful of wilderness areas in the northern Rocky Mountains of the USA. As a result, the climate and fire regime spaces of this sample of studies are not entirely representative of wilderness areas in general, and certainly not of broader geographic areas. We identify several wilderness areas that have experienced wildfire but few or no studies—under-represented areas that offer the possibility for future research to help expand the geographic, climate, and fire spaces of wilderness fire science. Finally, we urge continued research in wilderness areas that deepens our understanding of the past and potential role of cultural and Indigenous burning, the impacts of changing climate and fire regimes on ecosystems and landscape processes, and how to increase support for wilderness fire management and prescribed fire.

Availability of data and materials

The datasets and scripts generated and analyzed during the current study are available in the Zenodo repository, https://doi.org/10.5281/zenodo.6326355 .

Abatzoglou, J.T., S.Z. Dobrowski, S.A. Parks, and K.C. Hegewisch. 2018. TerraClimate, a high-resolution global dataset of monthly climate and climatic water balance from 1958–2015. Scientific Data 5: 170191.

Article   PubMed   PubMed Central   Google Scholar  

Agee, J.K. 1999. Fire effects on landscape fragmentation in interior West forests . Ledien, The Netherlands: Koninklijke Brill NV.

Google Scholar  

Agee, J.K. 2000. Wilderness fire science: a state of the knowledge review, in: Wilderness Ecosystems, Threats, and Management., Wilderness Science in a Time of Change Conferenc. USDA For. Serv. Gen. Tech. Rep. RMRS-P-15-VOL-5, pp. 5–22.

Arkle, R.S., D.S. Pilliod, and K. Strickler. 2010. Fire, flow and dynamic equilibrium in stream macroinvertebrate communities. Freshwater Biology 55: 299–314. https://doi.org/10.1111/j.1365-2427.2009.02275.x .

Article   Google Scholar  

Barnett, K., C. Miller, and T.J. Venn. 2016. Using Risk Analysis to Reveal Opportunities for the Management of Unplanned Ignitions in Wilderness. Journal of Forestry 114: 610–618. https://doi.org/10.5849/jof.15-111 .

Beckage, B., W.J. Platt, M.G. Slocum, and B. Panko. 2003. Influence of the El Niño Southern oscillation on fire regimes in the Florida everglades. Ecology 84: 3124–3130. https://doi.org/10.1890/02-0183 .

Belote, T., M. Dietz, and G. Aplet. 2015. Allocating Untreated “Controls” in the national wilderness preservation system as a climate adaptation strategy: a case study from the flathead national forest, Montana. Northwest Science 89: 239–254. https://doi.org/10.3955/046.089.0311 .

Berkey, J.K., R.T. Belote, C.T. Maher, and A.J. Larson. 2021a. Structural diversity and development in active fire regime mixed-conifer forests. Forest Ecology and Management 479: 118548.

Berkey, J.K., Miller, C., Larson, A.J. 2021b. A history of wilderness fire management in the Northern Rockies. Gen. Tech. Rep. RMRS-GTR-428. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. 88 p. https://doi.org/10.2737/RMRS-GTR-428 . 428. https://doi.org/10.2737/RMRS-GTR-428

Boisramé, G.F., S.E. Thompson, C. Tague, and S.L. Stephens. 2019. Restoring a natural fire regime alters the water balance of a Sierra Nevada catchment. Water Resources Research 55: 5751–5769.

Borrie, W., S. McCool, and J. Whitmore. 2006. Wildland fire effects on visits and visitors to the bob marshal wilderness complex. International Journal of Wilderness 12: 32–38.

Boxall, P.C., D.O. Watson, and J. Englin. 1996. Backcountry recreationists’ valuation of forest and park management features in wilderness parks of the western Canadian Shield. Can. J. For. Res. 26: 982–990. https://doi.org/10.1139/x26-108 .

Bozek, M.A., and M.K. Young. 1994. Fish mortality resulting from delayed-effects of fire in the greater yellowstone ecosystem. Great Basin Naturalist 54: 6.

Brown, R.N.K., Rosenberger, R.S., Kline, J.D., Hall, T.E., Needham, M.D. 2008. Visitor Preferences for Managing Wilderness Recreation after Wildfire. Journal of Forestry 8.

Bureau of Land Management, 2012. Management of Designated Wilderness Areas (Public), 6340.

Calkin, D.E., K.M. Gebert, J.G. Jones, and R.P. Neilson. 2005. Forest service large fire area burned and suppression expenditure trends, 1970–2002. Journal of Forestry. 103 (4): 179–183.

Collins, B.M., J.D. Miller, A.E. Thode, M. Kelly, J.W. van Wagtendonk, and S.L. Stephens. 2009. Interactions among wildland fires in a long-established Sierra Nevada natural fire area. Ecosystems 12: 114–128. https://doi.org/10.1007/s10021-008-9211-7 .

Coop, J.D., S.A. Parks, C.S. Stevens-Rumann, S.D. Crausbay, P.E. Higuera, M.D. Hurteau, A. Tepley, E. Whitman, T. Assal, B.M. Collins, K.T. Davis, S. Dobrowski, D.A. Falk, P.J. Fornwalt, P.Z. Fulé, B.J. Harvey, V.R. Kane, C.E. Littlefield, E.Q. Margolis, M. North, M.-A. Parisien, S. Prichard, and K.C. Rodman. 2020. Wildfire-driven forest conversion in Western North American landscapes. BioScience 70: 659–673. https://doi.org/10.1093/biosci/biaa061 .

Donaghy Cannon, M.M. 2013. The Use of Burned Forest By Breeding Birds Following the Biscuit Fire in Southwest Oregon (Ph.D.). Oregon State University, United States -- Oregon.

Dudley, N. 2013. Guidelines for applying protected area management categories including IUCN WCPA best practice guidance on recognising protected areas and assigning management categories and governance types. IUCN.

Dvorak, R.G., Small, E.D. 2011. Visitor attitudes towards fire and wind disturbances in wilderness. International Journal of Wilderness. 21(2): 27-31, 36. 21, 27-31,36.

Englin, J., Holmes, T.P., Lutz, J. 2008. Wildfire and the economic value of wilderness recreation. The Economics of Forest Disturbances: Wildfires, Storms, and Invasive Species, 191-208.

Fisher, A.H. 1997. The 1932 handshake agreement: Yakama Indian treaty rights and forest service policy in the Pacific Northwest. Western Historical Quarterly 28: 186–217. https://doi.org/10.2307/970893 .

Franklin, J.F., T.A. Spies, R.V. Pelt, A.B. Carey, D.A. Thornburgh, D.R. Berg, D.B. Lindenmayer, M.E. Harmon, W.S. Keeton, D.C. Shaw, K. Bible, and J. Chen. 2002. Disturbances and structural development of natural forest ecosystems with silvicultural implications, using Douglas-fir forests as an example. Forest Ecology and Management, Forest Ecology in the next Millennium : Putting the long view into Practice 155: 399–423. https://doi.org/10.1016/S0378-1127(01)00575-8 .

Franklin, J.F. 1987. Scientific use of wilderness, in: Lucas, R. C, Ed. Proceedings, national wilderness research conference: issues, State-of-knowledge, future directions., Gen. Tech. Rep. Ogden, UT, pp. 42–46.

Frelich, L.E. 2017. Wildland fire: understanding and maintaining an ecological baseline. Curr Forestry Rep 3: 188–201. https://doi.org/10.1007/s40725-017-0062-3 .

Frelich, L.E., and P.B. Reich. 2009. Wilderness conservation in an era of global warming and invasive species: a case study from Minnesota’s boundary waters canoe area wilderness. Naar 29: 385–393.

Fulé, P.Z., and D.C. Laughlin. 2007. Wildland fire effects on forest structure over an altitudinal gradient, Grand Canyon National Park, USA. Journal of Applied Ecology 44: 136–146. https://doi.org/10.1111/j.1365-2664.2006.01254.x .

Gorelick, N., M. Hancher, M. Dixon, S. Ilyushchenko, D. Thau, and R. Moore. 2017. Google Earth Engine: Planetary-scale geospatial analysis for everyone. Remote Sensing of Environment 202: 18–27. https://doi.org/10.1016/j.rse.2017.06.031 .

Habeck, J.R., and R.W. Mutch. 1973. Fire-dependent forests in the Northern Rocky Mountains. Quaternary Research, The Ecological Role of Fire in Natural Conifer Forests of Western and Northern America 3: 408–424. https://doi.org/10.1016/0033-5894(73)90006-9 .

Haire, S., K. McGarigal, and C. Miller. 2013. Wilderness shapes contemporary fire size distributions across landscapes of the western United States. Ecosphere 4 (1): 1–20. https://doi.org/10.1890/ES12-00257.1 .

Heinselman, M.L. 1973. Fire in the virgin forests of the Boundary Waters Canoe Area, Minnesota. Quaternary Research, The Ecological Role of Fire in Natural Conifer Forests of Western and Northern America 3: 329–382. https://doi.org/10.1016/0033-5894(73)90003-3 .

Heinselman, M.L. 1970. Restoring fire to the ecosystems of the Boundary Waters Canoe Area, Minnesota, and to similar wilderness areas., in: Proceedings: 10th Tall Timbers Fire Ecology Conference 1970.

Holden, Z.A., P. Morgan, M.G. Rollins, and R.G. Wright. 2006. Ponderosa pine snag densities following multiple fires in the Gila Wilderness, New Mexico. Forest Ecology and Management 221: 140–146. https://doi.org/10.1016/j.foreco.2005.09.014 .

Hopkins, T., A.J. Larson, and R.T. Belote. 2014. Contrasting effects of wildfire and ecological restoration in old-growth Western Larch forests. Forest Science 60: 1005–1013. https://doi.org/10.5849/forsci.13-088 .

Jackson, B.K., and S.M.P. Sullivan. 2009. Influence of wildfire severity on riparian plant community heterogeneity in an Idaho, USA wilderness. Forest Ecology and Management 259: 24–32. https://doi.org/10.1016/j.foreco.2009.09.036 .

Jackson, B.K., S.M.P. Sullivan, and R.L. Malison. 2012. Wildfire severity mediates fluxes of plant material and terrestrial invertebrates to mountain streams. Forest Ecology and Management 278: 27–34. https://doi.org/10.1016/j.foreco.2012.04.033 .

Jacobs, G.R., R.F. Thurow, J.M. Buffington, D.J. Isaak, and S.J. Wenger. 2021. Climate, fire regime, geomorphology, and conspecifics influence the spatial distribution of chinook salmon redds. Transactions of the American Fisheries Society 150: 8–23. https://doi.org/10.1002/tafs.10270 .

Jaffe, D.A., S.M. O’Neill, N.K. Larkin, A.L. Holder, D.L. Peterson, J.E. Halofsky, and A.G. Rappold. 2020. Wildfire and prescribed burning impacts on air quality in the United States. Journal of the Air & Waste Management Association 70: 583–615. https://doi.org/10.1080/10962247.2020.1749731 .

Article   CAS   Google Scholar  

Jain, P., D. Castellanos-Acuna, S.C.P. Coogan, J.T. Abatzoglou, and M.D. Flannigan. 2022. Observed increases in extreme fire weather driven by atmospheric humidity and temperature. Nat. Clim. Chang. 12: 63–70. https://doi.org/10.1038/s41558-021-01224-1 .

Kane, V.R., J.A. Lutz, S.L. Roberts, D.F. Smith, R.J. McGaughey, N.A. Povak, and M.L. Brooks. 2013. Landscape-scale effects of fire severity on mixed-conifer and red fir forest structure in Yosemite National Park. Forest Ecology and Management 287: 17–31. https://doi.org/10.1016/j.foreco.2012.08.044 .

Kay, C.E. 2000. Native Burning in Western North America: Implications for Hardwood Forest Management. In: Yaussy, Daniel A., comp. 2000. Proceedings: workshop on fire, people, and the central hardwoods landscape; 2000 March 12-14; Richmond, KY. Gen. Tech. Rep. NE-274. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, Northeastern Research Station: 19-27. 274.

Keane, R.E., S. Arno, and L.J. Dickinson. 2006. The complexity of managing fire-dependent ecosystems in wilderness: relict ponderosa pine in the bob marshall wilderness. Ecological Restoration 24: 71–78. https://doi.org/10.3368/er.24.2.71 .

Kilgore, B.M. 1973. The ecological role of fire in sierran conifer forests: its application to national park management. Quaternary Research 3: 496–513. https://doi.org/10.1016/0033-5894(73)90010-0 .

Kilgore, B.M. 1987. The role of fire in wilderness: a state-of-knowledge review, Proceedings of the national wilderness research conference: Issues, state-of-knowledge, future directions, compiled by RC Lucas. Intermountain Research Station.

Kimmerer, R.W., and F.K. Lake. 2001. The role of indigenous burning in land management. Journal of Forestry 99: 36–41. https://doi.org/10.1093/jof/99.11.36 .

Kipfmueller, K.F., E.R. Larson, L.B. Johnson, and E.A. Schneider. 2021. Human augmentation of historical red pine fire regimes in the boundary waters Canoe area wilderness. Ecosphere 12: e03673. https://doi.org/10.1002/ecs2.3673 .

Kleindl, W.J., M.C. Rains, L.A. Marshall, and F.R. Hauer. 2015. Fire and flood expand the floodplain shifting habitat mosaic concept. Freshwater Science 34: 1366–1382. https://doi.org/10.1086/684016 .

Knotek, K., A.E. Watson, W.T. Borrie, J.G. Whitmore, and D. Turner. 2008. Recreation visitor attitudes towards management-ignited prescribed fires in the bob marshall wilderness complex, Montana. Journal of Leisure Research 40: 608–618. https://doi.org/10.1080/00222216.2008.11950155 .

Kuuluvainen, T., P. Angelstam, L. Frelich, K. Jõgiste, M. Koivula, Y. Kubota, B. Lafleur, and E. Macdonald. 2021. Natural disturbance-based forest management: moving beyond retention and continuous-cover forestry. Frontiers in Forests and Global Change 4: 629020.

Landres, P. 2010. A framework to evaluate proposals for scientific activities in wilderness. Gen. Tech. Rep. RMRS-GTR-234. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. 74 p. 234. https://doi.org/10.2737/RMRS-GTR-234

Larson, A.J., R.T. Belote, C.A. Cansler, S.A. Parks, and M.S. Dietz. 2013. Latent resilience in ponderosa pine forest: effects of resumed frequent fire. Ecological Applications 23: 1243–1249. https://doi.org/10.1890/13-0066.1 .

Article   PubMed   Google Scholar  

Larson, E.R., K.F. Kipfmueller, and L.B. Johnson. 2021. People, fire, and pine: linking human agency and landscape in the boundary waters canoe area wilderness and beyond. Annals of the American Association of Geographers 111: 1–25. https://doi.org/10.1080/24694452.2020.1768042 .

Lawhon, J. 2011. Prescribed Fire in Wilderness: Nature or Nurture? Rethinking protected areas in a changing world: proceedings of the 2011 George Wright Society Conference on Parks, Protected Areas and Cultural Sites 177–184.

Leopold, A.S., S.A. Cain, C.M. Cottam, I.N. Gabrielson, and T.L. Kimball. 1963. Wildlife Management in the National Parks . National Park Service: U.S.

Lotan, J.E., Kilgore, B.M., Fischer, W.C., Mutch, R.M. 1985. Proceedings—symposium and workshop on wilderness fire; 1983 November 15-18; Missoula, MT. Gen. Tech. Rep. INT-GTR-182. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Forest and Range Experiment Station. 434 p. 182.

Love, T.G., Watson, A.E. 1992. Effects of the Gates Park Fire on recreation choices (No. INT-RN-402). U.S. Department of Agriculture, Forest Service, Intermountain Research Station, Ogden, UT. https://doi.org/10.2737/INT-RN-402

Malison, R.L., and C.V. Baxter. 2010. Effects of wildfire of varying severity on benthic stream insect assemblages and emergence. Journal of the North American Benthological Society 29: 1324–1338. https://doi.org/10.1899/09-022.1 .

Marcus, W.A., J. Rasmussen, and M.A. Fonstad. 2011. Response of the fluvial wood system to fire and floods in Northern Yellowstone. Annals of the Association of American Geographers 101: 21–44. https://doi.org/10.1080/00045608.2010.539154 .

McCool, S.F., Stankey, G.H. 1986. Visitor Attitudes Toward Wilderness Fire Management Policy -- 1971-84. United States Department of Agriculture, Forest Service Research Paper Int-357.

Miller, C. 2012. The hidden consequences of fire suppression. Park Science 28: 75–80.

Miller, C., and G.H. Aplet. 2016. Progress in wilderness fire science: embracing complexity. Journal of Forestry 114: 373–383. https://doi.org/10.5849/jof.15-008 .

Miller, C., J. Abatzoglou, T. Brown, and A.D. Syphard. 2011. Wilderness Fire Management in a Changing Environment. In The Landscape Ecology of Fire, Ecological Studies , ed. D. McKenzie, C. Miller, and D.A. Falk, 269–294. Netherlands, Dordrecht: Springer.

Chapter   Google Scholar  

Miller, C. 2003. Wildland fire use: a Wilderness perspective on fuel management. USDA forest service proceedings RMRS-P-29.

Minshall, G.W., C.T. Robinson, D.E. Lawrence, D.A. Andrews, and J.T. Brock. 2001. Benthic macroinvertebrate assemblages in five central Idaho (USA) streams over a 10-year period following disturbance by wildfire. Int. J. Wildland Fire 10: 201. https://doi.org/10.1071/WF01018 .

Morgan, P., E.K. Heyerdahl, C. Miller, A.M. Wilson, and C.E. Gibson. 2014. Northern Rockies Pyrogeography: an example of fire Atlas utility. Fire Ecology 10: 14–30. https://doi.org/10.4996/fireecology.1001014 .

National Park Service, 2006. Management Policies.

Nesmith, J.C.B., A.C. Caprio, A.H. Pfaff, T.W. McGinnis, and J.E. Keeley. 2011. A comparison of effects from prescribed fires and wildfires managed for resource objectives in Sequoia and Kings Canyon National Parks. Forest Ecology and Management 261: 1275–1282. https://doi.org/10.1016/j.foreco.2011.01.006 .

Ostlund, L., B. Keane, S. Arno, and R. Andersson. 2005. Culturally scarred trees in the Bob Marshall Wilderness, Montana, USA - Interpreting Native American historical forest use in a wilderness area. Natural Areas Journal. 25 (4): 315–325.

Parks, S.A., C. Miller, C.R. Nelson, and Z.A. Holden. 2014. Previous fires moderate burn severity of subsequent wildland fires in two large Western US wilderness areas. Ecosystems 17: 29–42. https://doi.org/10.1007/s10021-013-9704-x .

Parks, S.A., L.M. Holsinger, C. Miller, and C.R. Nelson. 2015. Wildland fire as a self-regulating mechanism: the role of previous burns and weather in limiting fire progression. Ecological Applications 25: 1478–1492. https://doi.org/10.1890/14-1430.1 .

Parks, S.A., C. Miller, L.M. Holsinger, L.S. Baggett, and B.J. Bird. 2016. Wildland fire limits subsequent fire occurrence. Int. J. Wildland Fire 25: 182. https://doi.org/10.1071/WF15107 .

Parsons, D.J., Landres, P.B., Miller, C. 2003. Wildland fire use: the dilemma of managing and restoring natural fire and fuels in United States wilderness. In: Galley, K.E.M.; Klinger, R.C.; Sugihara, N.G., eds. Proceedings of Fire Conference 2000: the First National Congress on Fire Ecology, Prevention, and Management. Miscellaneous Publication ; no. 13.Tallahassee, FL : Tall Timbers Research Station: 19-26.

Parsons, D.J. 2000. The challenge of restoring natural fire to wilderness. In: Cole, David N.; McCool, Stephen F.; Borrie, William T.; O’Loughlin, Jennifer, comps. 2000. Wilderness science in a time of change conference-Volume 5: Wilderness ecosystems, threats, and management; 1999 May 23–27; Missoula, MT. Proceedings RMRS-P-15-VOL-5. Ogden, UT: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. p. 276-282 015.

Parsons, D.J. 2007. An outside assessment of wilderness research in the forest service. International Journal of Wilderness 13 (3): 34–35, 39.

Pawlikowski, N.C., M. Coppoletta, E. Knapp, and A.H. Taylor. 2019. Spatial dynamics of tree group and gap structure in an old-growth ponderosa pine-California black oak forest burned by repeated wildfires. Forest Ecology and Management 434: 289–302. https://doi.org/10.1016/j.foreco.2018.12.016 .

Peterson, G.D. 2002. Contagious disturbance, ecological memory, and the emergence of landscape pattern. Ecosystems 5: 329–338. https://doi.org/10.1007/s10021-001-0077-1 .

Pullin, A.S., and G.B. Stewart. 2006. Guidelines for systematic review in conservation and environmental management. Conservation Biology 20: 1647–1656. https://doi.org/10.1111/j.1523-1739.2006.00485.x .

Quadrennial Fire Review. 2014. US Department of Agriculture Forest Service Fire and Aviation Management and Department of the Interior Office of Wildland Fire, Washington, DC.

Reilly, M.J., M.C. Wimberly, and C.L. Newell. 2006. Wildfire effects on plant species richness at multiple spatial scales in forest communities of the southern Appalachians. J Ecology 94: 118–130. https://doi.org/10.1111/j.1365-2745.2005.01055.x .

Robinne, F.-N., Burns, J., Kant, P., Groot, B., Kleine, M., Flannigan, M., Albar, I., Amissah, L., Bowman, D., Charlton, V., Cochrane, M., Field, R., Goldammer, J., Khan, M., Kornexl, W., Krasovskiy, A., Mccaffrey, S., Mitchell, A., Mohanty, A.K., Ellison, D. 2019. Global Fire Challenges in a Warming World. World Bank.

Robinson, C.T., U. Uehlinger, and G.W. Minshall. 2005. Functional characteristics of wilderness streams twenty years following wildfire. Western North American Naturalist 65: 1–10.

Rollins, M.G. 2009. LANDFIRE: a nationally consistent vegetation, wildland fire, and fuel assessment. International Journal of Wildland Fire 18: 235. https://doi.org/10.1071/WF08088 .

Roos, C.I., T.W. Swetnam, T.J. Ferguson, M.J. Liebmann, R.A. Loehman, J.R. Welch, E.Q. Margolis, C.H. Guiterman, W.C. Hockaday, M.J. Aiuvalasit, J. Battillo, J. Farella, and C.A. Kiahtipes. 2021. Native American fire management at an ancient wildland–urban interface in the Southwest United States. Proceedings of the National Academy of Sciences 118: e2018733118.

Rothman, H. 2007. Blazing Heritage: A History of Wildland Fire in the National Parks . USA: Oxford University Press.

Book   Google Scholar  

Rugenski, A.T., and G.W. Minshall. 2014. Climate-moderated responses to wildfire by macroinvertebrates and basal food resources in montane wilderness streams. Ecosphere 5: art25. https://doi.org/10.1890/ES13-00236.1 .

Ruiz, P.L., J.P. Sah, M.S. Ross, and A.A. Spitzig. 2013. Tree Island Response to Fire and Flooding in the Short-Hydroperiod Marl Prairie Grasslands of the Florida Everglades, USA. Fire Ecology 9: 38–54.

Sánchez, J.J., K. Baerenklau, and A. González-Cabán. 2016. Valuing hypothetical wildfire impacts with a Kuhn-Tucker model of recreation demand. Forest Policy and Economics 71: 63–70. https://doi.org/10.1016/j.forpol.2015.08.001 .

Schoennagel, T., J.K. Balch, H. Brenkert-Smith, P.E. Dennison, B.J. Harvey, M.A. Krawchuk, N. Mietkiewicz, P. Morgan, M.A. Moritz, R. Rasker, M.G. Turner, and C. Whitlock. 2017. Adapt to more wildfire in western North American forests as climate changes. PNAS 114: 4582–4590. https://doi.org/10.1073/pnas.1617464114 .

Article   CAS   PubMed   PubMed Central   Google Scholar  

Schroeder, S.L., and I.E. Schneider. 2010. Wildland fire and the wilderness visitor experience. International Journal of Wilderness 16 (1): 20–25.

Schultz, C.A., Huber-Stearns, H., McCaffrey, S., Quirke, D., Ricco, G., Moseley, C. 2018. Prescribed fire policy barriers and opportunities : a diversity of challenges and strategies across the west (Working Paper). Ecosystem Workforce Program, Institute for a Sustainable Environment, University of Oregon.

Scott, J.H., D.J. Helmbrecht, S.A. Parks, and C. Miller. 2012. Quantifying the Threat of Unsuppressed Wildfires Reaching the Adjacent Wildland-Urban Interface on the Bridger-Teton National Forest, Wyoming, USA. Fire Ecology 8: 125–142. https://doi.org/10.4996/fireecology.0802125 .

Seielstad, C. 2015. Reconsidering wildland fire use: Perspectives from the Northern Rockies. In: Keane, Robert E.; Jolly, Matt; Parsons, Russell; Riley, Karin. Proceedings of the large wildland fires conference; May 19-23, 2014; Missoula, MT. Proc. RMRS-P-73. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. p. 207-212. 73, 207–212.

Smith, D. 2014. From research to policy: the white cap wilderness fire study. Forest History Today. Spring/Fall 4–12: 4–12.

Smith, R.J., and A.N. Gray. 2021. Strategic monitoring informs wilderness management and socioecological benefits. Conservation Science and Practice 3: e482. https://doi.org/10.1111/csp2.482 .

Spencer, C.N., K.O. Gabel, and F.R. Hauer. 2003. Wildfire effects on stream food webs and nutrient dynamics in Glacier National Park, USA. Forest Ecology and Management, The Effect of Wildland Fire on Aquatic Ecosystems in the Western USA 178: 141–153. https://doi.org/10.1016/S0378-1127(03)00058-6 .

Suffling, R., A. Grant, and R. Feick. 2008. Modeling prescribed burns to serve as regional firebreaks to allow wildfire activity in protected areas. Forest Ecology and Management 256: 1815–1824. https://doi.org/10.1016/j.foreco.2008.06.043 .

Talley, S.N., and J.R. Griffin. 1980. Fire ecology of a montane pine forest, Junipero Serra peak, California. Madroño 27: 49–60.

Tanner, S., Lupi, F., Garnache, C., Tanner, S., Lupi, F., Garnache, C., 2022. Estimating visitor preferences for recreation sites in wildfire prone areas. Int. J. Wildland Fire 31, 871–885. https://doi.org/10.1071/WF21133 .

Taylor, A.H. 2010. Fire disturbance and forest structure in an old-growth Pinus ponderosa forest, southern Cascades, USA. Journal of Vegetation Science 21: 561–572. https://doi.org/10.1111/j.1654-1103.2009.01164.x .

Teske, C.C., C.A. Seielstad, and L.P. Queen. 2012. Characterizing Fire-on-Fire Interactions in Three Large Wilderness Areas. Fire Ecology 8: 82–106. https://doi.org/10.4996/fireecology.0802082 .

Thompson, J.R., and T.A. Spies. 2009. Vegetation and weather explain variation in crown damage within a large mixed-severity wildfire. Forest Ecology and Management 258: 1684–1694. https://doi.org/10.1016/j.foreco.2009.07.031 .

Trauernicht, C., B.P. Murphy, N. Tangalin, and D.M.J.S. Bowman. 2013. Cultural legacies, fire ecology, and environmental change in the Stone Country of Arnhem Land and Kakadu National Park, Australia. Ecol Evol 3: 286–297. https://doi.org/10.1002/ece3.460 .

Turner, M.G. 1989. Landscape ecology: the effect of pattern on process. Annual Review of Ecology and Systematics 20: 171–197.

U.S. Fish and Wildlife Service. 2008. Wilderness Administration and Resource Stewardship, 610 FW 2.

U.S. Forest Service. 2007. Fire Management in Wilderness, Forest Service Manual (FSM) 2320.

Watson, A., S. Martin, N. Christensen, G. Fauth, and D. Williams. 2015. The relationship between perceptions of wilderness character and attitudes toward management intervention to adapt biophysical resources to a changing climate and nature restoration at sequoia and kings canyon national parks. Environmental Management 56: 653–663. https://doi.org/10.1007/s00267-015-0519-8 .

Watson, A., Matt, R., Knotek, K., Williams, D., Yung, L. 2011. Traditional wisdom – protecting wilderness as a cultural landscape. Ecology and Society 16. https://doi.org/10.5751/ES-04003-160136

Wilderness Act 16 U.S. Code § 1131 [WWW Document], 1964. URL https://uscode.house.gov/view.xhtml?path=/prelim@title16/chapter23&edition=prelim (accessed 4.15.21).

Williamson, M.A. 2007. Factors in United States Forest Service district rangers’ decision to manage a fire for resource benefit. Int. J. Wildland Fire 16: 755–762. https://doi.org/10.1071/WF06019 .

Wright, H.E. 1974. Landscape development, forest fires, and wilderness management. Science 186: 487–495. https://doi.org/10.1126/science.186.4163.487 .

Van Wagtendonk, J.W. 2007. The history and evolution of wildland fire use. Fire Ecology 3: 3–17.

van Wagtendonk, J.W., K.A. van Wagtendonk, and A.E. Thode. 2012. Factors associated with the severity of intersecting fires in Yosemite National Park, California, USA. Fire Ecology 8: 11–31.

Vequist, G. 2007. Ecological restoration of degraded wilderness ecosystems: Removing exotic plants and introducing prescribed fire to restore natural diversity in two National Park wilderness areas. In: Watson, Alan; Sproull, Janet; Dean, Liese, comps. Science and stewardship to protect and sustain wilderness values: Eighth World Wilderness Congress symposium; September 30-October 6, 2005; Anchorage, AK. Proceedings RMRS-P-49. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. p. 506-509 049.

Zimmerman, T., Frary, M., Crook, S., Fay, B., Koppenol, P., Lasko, R. 2006. Wildland Fire Use - Challenges Associated With Program Management Across Multiple Ownerships and Land Use Situations. In: Andrews, Patricia L.; Butler, Bret W., comps. 2006. Fuels Management-How to Measure Success: Conference Proceedings. 28-30 March 2006; Portland, OR. Proceedings RMRS-P-41. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. p. 47-58 041.

Download references

Acknowledgements

We thank Bethany Allen for assistance in the literature search, and Carol Miller for supporting this project at the earliest stages. This research was supported in part by the USDA Forest Service, Rocky Mountain Research Station, Aldo Leopold Wilderness Research Institute. The findings and conclusions in this publication are those of the authors and should not be construed to represent any official USDA or U.S. Government determination or policy.

Funding support provided by US Forest Service Rocky Mountain Research Station agreement number 17-JV-11221639-092. This material is based upon work supported by the National Science Foundation under Grant No. 1745048.

Author information

Authors and affiliations.

Wilderness Institute, University of Montana, W.A. Franke College of Forestry and Conservation, 32 Campus Drive, Missoula, MT, 59812, USA

Mark R. Kreider, Melissa R. Jaffe & Andrew J. Larson

Montana Department of Natural Resources and Conservation, 1401 27th Avenue, Missoula, MT, 59804, USA

Julia K. Berkey

Aldo Leopold Wilderness Research Institute, Rocky Mountain Research Station, USDA Forest Service, 790 East Beckwith Avenue, Missoula, MT, 59801, USA

Sean A. Parks

You can also search for this author in PubMed   Google Scholar

Contributions

Mark R. Kreider: Investigation, Formal analysis, Data curation, Writing—original draft, Writing – review & editing, Visualization. Melissa R. Jaffe: Investigation, Writing – review & editing. Julia K. Berkey: Methodology, Writing – review & editing. Sean A. Parks: Investigation, Writing – review & editing. Andrew J. Larson: Conceptualization, Methodology, Investigation, Writing – original draft, Writing – review & editing, Supervision, Project administration, Funding acquisition.

Corresponding author

Correspondence to Mark R. Kreider .

Ethics declarations

Ethics approval and consent to participate.

Not applicable.

Consent for publication

Competing interests.

The authors declare that they have no competing interests.

Additional information

Publisher’s note.

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

Supplementary Information

Additional file 1., additional file 2., additional file 3., rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Kreider, M.R., Jaffe, M.R., Berkey, J.K. et al. The scientific value of fire in wilderness. fire ecol 19 , 36 (2023). https://doi.org/10.1186/s42408-023-00195-2

Download citation

Received : 03 March 2022

Accepted : 17 May 2023

Published : 14 June 2023

DOI : https://doi.org/10.1186/s42408-023-00195-2

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Climate change
  • Prescribed burning
  • Resource objective fire
  • Wilderness for science
  • Wilderness fire management
  • Wildland fire use

research paper on fire ecology

IMAGES

  1. Fire Ecology Journal

    research paper on fire ecology

  2. PPT

    research paper on fire ecology

  3. Introduction to Fire Ecology (Lesson 1 of 6)

    research paper on fire ecology

  4. (PDF) "Fire Ecology Across Boundaries": the international congress

    research paper on fire ecology

  5. (PDF) The Landscape Ecology of Fire

    research paper on fire ecology

  6. Revision for Fire and Ecology SCENV3110

    research paper on fire ecology

COMMENTS

  1. Fire as a fundamental ecological process: Research advances and frontiers

    Our central goal in this paper is to advance knowledge of fire ecology. As the magnitude of fire ecology research increases, it becomes increasingly important to identify priorities in understanding the ecological and evolutionary implications of changing fire activity. Here, we attempt to synthesize the major areas of fire ecology research.

  2. Home

    Fire Ecology, in partnership with Research Square, now offers In Review: a journal-integrated preprint service. Authors choosing this free optional service will be able to: ... These papers represent diverse contributions relating fire to flora, fauna, water, humans, and more. The multiple submission types are the products of the diverse ...

  3. Changing wildfire, changing forests: the effects of ...

    Background Wildfires in the Pacific Northwest (Washington, Oregon, Idaho, and western Montana, USA) have been immense in recent years, capturing the attention of resource managers, fire scientists, and the general public. This paper synthesizes understanding of the potential effects of changing climate and fire regimes on Pacific Northwest forests, including effects on disturbance and stress ...

  4. Articles

    Globally, prescribed fire political interest and practice has been rekindled following recent devastating wildfire seasons. This phenomenon was especially acute in areas with dual wildfire and forest health cr... Kate M. Wilkin, Amanda M. Stasiewicz and Susan D. Kocher. Fire Ecology 2024 20 :48.

  5. Fire ecology

    Fire ecology also studies the effect of anthropogenic change and management on the incidence and effects of wildfires. ... Research Highlights 28 Apr 2023 Nature Reviews Earth & Environment ...

  6. Fire Ecology Journal

    Fire Ecology is an international scientific journal supported by the Association for Fire Ecology and published by SpringerOpen. You can freely access and submit articles on the journal website. Also, make sure to follow @SpringerEcology and @fireecology for the latest posts, or search #SNFECO. Read Articles in Fire Ecology.

  7. Fire ecology for the 21st century: Conserving biodiversity in the age

    Five core areas of research in megafire ecology are summarised below: Identifying the drivers of megafire in search of options for managing fire size, intensity, and severity, including the relative and interacting roles of climate and land use (Khorshidi et al., 2020; Stephens et al., 2014).Of increasing interest is the potential for cultural fire to play a role in diminishing the threat of ...

  8. Fire as a fundamental ecological process: Research advances and

    Journal of Ecology publishes original research on all aspects of the plant ecology (including algae), in both aquatic and terrestrial ecosystems. Abstract Fire is a powerful ecological and evolutionary force that regulates organismal traits, population sizes, species interactions, community composition, carbon and nutrient cycling and ecosys...

  9. The eco‐evolutionary role of fire in shaping terrestrial ecosystems

    Here, we synthesize these contributions to the Special Feature on fire as a dynamic ecological and evolutionary force and place them in a broader context of fire research. 1 PLANT RESPONSE TO FIRE. Understanding how species evolve to persist in novel fire regimes requires fundamental knowledge of the feedbacks involving climate, fire and ...

  10. Fire ecology

    Read the latest Research articles in Fire ecology from Scientific Reports

  11. Influences of wildfire on the forest ecosystem and climate change: A

    The influence of a forest fire differs based on the vegetational composition of the forest neighborhood, the quantity of fuel demand, the frequency, extent, and the resulting fire ascend interval associated with that forest (Littell et al., 2016).Around the world, forest fires consume around 4.5 million Km 2 each year (Mansoor et al., 2022). Fig. 2 denotes that the forest lost to fires ...

  12. A global synthesis of fire effects on ecosystem services of forests and

    Frontiers in Ecology and the Environment is an environmental science journal publishing articles that address current and emerging ecological and environmental issues. ... we included papers containing "fire" or other similar terms in the title (eg "wildfire", "burn", and so forth). ... Current status of global fire research on ...

  13. Global patterns and drivers of post-fire vegetation ...

    Here we present the global estimates of post-fire vegetation productivity recovery from 2004 to 2021 using gross primary productivity observations and related proxies at a spatial resolution of 10 ...

  14. 30519 PDFs

    Explore the latest full-text research PDFs, articles, conference papers, preprints and more on FIRE ECOLOGY. Find methods information, sources, references or conduct a literature review on FIRE ...

  15. Frontiers in Fire Ecology

    Frontiers in Fire Ecology. Edited by: Karin L. Riley, Leda Kobziar, Andrew T. Hudak. Association for Fire Ecology celebrated its 20th anniversary in 2017 at the 7th International Fire Ecology and Management Congress. More than 350 research presentations highlighted advances in fire ecology and fire management science, policy, and practice.

  16. Fire

    Ecosystems are frequently disturbed by fires that have an important impact on the soil environment and the composition of soil organisms. In order to provide a baseline for the current research and identify trends on the effects of wildland fire on soil environment and biological changes, the available literature was identified from the Web of Science database, covering the period from 1998 ...

  17. Vegetation's influence on fire behavior goes beyond ...

    Many fire ecology studies are focused on fire effects on vegetation as framed by a particular ecosystem's fire regime. This creates an often-circular logic, since the term "fire regime" is defined by the broad-scale and long-term effects of fire frequency and intensity on overall ecosystem response, post-fire recovery, and successional patterns (Archibald et al. 2013).

  18. Evolutionary fire ecology: An historical account and future directions

    Among recent papers that have been focused on fire melanism, a key example is the study of natural selection to postfire conditions in grasshoppers (Forsman et al. 2011); apparently, no such studies have been done in vertebrates. Early evolutionary fire ecology: Flora

  19. (PDF) Impact of Forest Fire on Forest Ecosystem

    Impact of forest fire on forest ecosystem. J Agric Technol. 9 (1&2): 18-29. Fire is a common hazard in forest which plays an important role in shaping a community with its. ecosystem components ...

  20. Causes and Effects of Forest Fires in Tropical Rainforests: A

    The last two subjects include the effects of fire on ecological variables and are related to the research on soil and biomass and tree species. The topic related to soil and biomass was less frequently reported, while studies on the relationship between fire and tree species drew the attention of researchers in late 90s and the first decade of ...

  21. What Is Fire Ecology?

    Fire ecology is the study of fires and fire regimes in global forest, prairie, shrubland, chaparral, meadow, and savannah ecosystems. Fire ecologists study how changes in regime factors, like the frequency, severity, and extent of fires can provide favorable or damaging influences on ecosystems. They also study the interactions between fires ...

  22. Fire Ecology and Research

    Learn about Fire Research & Fire Ecology. Fire managers look to scientific research for guidance on how prescribed burns and lightning-started fires can can be used to benefit ecological systems. They examine historic fire regimes and past cultural practices, and use predictive modeling to study future fire effects under different climates and ...

  23. Introduction to fire ecology across USA forested ecosystems: Past

    In this book, we introduce the driving forces, historical patterns, and future management challenges of fire in forested ecoregions across the continental USA. Climate warming and decades of fire suppression or exclusion have altered historical fire regimes and threaten diversity of fire-adapted forest vegetation into the future.

  24. Environmental Science & Ecology Alum & Faculty Publish Peer-Reviewed Paper

    Jarrod Ludwig (BS/MS'24) and Dr. Jacques Rinchard published a paper in Journal of Great Lakes Research. This paper, based on Jarrod's SURP project from his undergraduate studies in 2022, provides insights into the reproductive spawning strategy of the deepwater sculpin through histological analysis.

  25. Assessing changes in global fire regimes

    Fire Ecology volume 20, ... we invited coauthors from the papers in the background review as well as referrals from workshop participants and all respondents who filled out the questionnaire. ... The 2019/2020 mega-fires exposed Australian ecosystems to an unprecedented extent of high-severity fire. Environmental Research Letters 16(4): 044029.

  26. Wildfires and Real Estate Values in California

    Wildfires have been a concern in California for decades. The intensity of these events has increased recently, with particularly large and destructive fire seasons between 2018 and 2021. Analysis shows that distance from high fire-risk zones had little impact on residential housing values in the past. However, that has changed since the late 2010s, coinciding with more extensive fire damage to ...

  27. Fact-checking warnings from Democrats about Project 2025 and ...

    Vice President Kamala Harris, the Democratic presidential nominee, has warned Americans about "Trump's Project 2025" agenda — even though former President Donald Trump doesn't claim the ...

  28. The scientific value of fire in wilderness

    Today, Indigenous and prescribed fire will likely play an important role in addressing the impacts of changing climate and fire regimes in protected areas, and increased research and discussion is vital (Larson et al. 2021 ). Our review identified over 220 scientific studies enabled by fire in wilderness.

  29. Full article: The ecology of Nkolola girls' initiation rites in the

    Nkolola ritual: definition and rationale. The term Nkolola is derived from lukolo, 'breast', inkolo, 'breasts', the manifest sign that a woman was ready for child bearing among precolonial Tonga society (Colson, Citation 2006, p. 159).The term 'Nkolola' is ambiguous as it is often used to refer to the girl under initiation; sometimes it refers to the ceremony itself; and at other ...