scispace - formally typeset
Search or ask a question
Posted ContentDOI

Modeling long-term fire impact on ecosystem characteristics and surface energy using a process-based vegetation–fire model SSiB4/TRIFFID-Fire v1.0

02 Dec 2020-Geoscientific Model Development (Copernicus GmbH)-Vol. 13, Iss: 12, pp 6029-6050
TL;DR: In this article, a fire model of intermediate complexity to a biophysical model with dynamic vegetation, SSiB4/TRIFFID (The Simplified Simple Biosphere Model coupled with the Top-down Representation of Interactive Foliage and Flora Including Dynamics Model), is used to simulate the burned area during 1948-2014.
Abstract: . Fire is one of the primary disturbances to the distribution and ecological properties of the world’s major biomes and can influence the surface fluxes and climate through vegetation-climate interactions. This study incorporates a fire model of intermediate complexity to a biophysical model with dynamic vegetation, SSiB4/TRIFFID (The Simplified Simple Biosphere Model coupled with the Top-down Representation of Interactive Foliage and Flora Including Dynamics Model). This new model, SSiB4/TRIFFID-Fire, updating fire impact on the terrestrial carbon cycle every 10 days, is then used to simulate the burned area during 1948–2014. The simulated global burned area in 2000–2014 is 471.9 Mha yr−1, close to the estimate, 478.1 Mha yr−1, in Global Fire Emission Database v4s (GFED4s) with a spatial correlation of 0.8. The SSiB4/TRIFFID-Fire reproduces temporal variations of the burned area at monthly to interannual scales. Specifically, it captures the observed decline trend in northern African savanna fire and accurately simulates the fire seasonality in most major fire regions. The simulated fire carbon emission is 2.19 Pg yr−1, slightly higher than the GFED4s (2.07 Pg yr−1). The SSiB4/TRIFFID-Fire is applied to assess long-term fire impact on ecosystem characteristics and surface energy budget by comparing model runs with and without fire (FIRE-ON minus FIRE-OFF). The FIRE-ON simulation reduces tree cover over 6.14 % of the global land surface, accompanied by a decrease in leaf area index and vegetation height by 0.13 m2 m−2 and 1.27 m, respectively. The surface albedo and sensible heat are reduced throughout the year, while latent heat flux decreases in the fire season but increases in the rainy season. Fire results in an increase in surface temperature over most fire regions.

Summary (4 min read)

1 Introduction

  • Many empirically determined parameters are included in the complicated process-based models, which leads to large uncertainties.
  • Moreover, current fire models have deficiencies in simulating the peak fire month, fire season length, and interannual variability, as reported by the Fire Model Intercomparison Project (FireMIP; Hantson et al., 2020; Li et al., 2019).
  • The SSiB4/TRIFFID-Fire model updates fire-induced carbon loss every 10 d, which has been rarely employed in current process-based fire models, and is used to provide a quantitative assessment of fire impact on ecosystem characteristics and surface energy at subseasonal, seasonal, interannual, and long-term scales.

2.1 Land and vegetation model

  • The Simplified Simple Biosphere Model (SSiB, Xue et al., 1991; Zhan et al., 2003) is a biophysical model which simulates fluxes of surface radiation, momentum, sensible and Geosci.
  • Latent heat, runoff, soil moisture, surface temperature, and vegetation gross and net primary productivity (GPP and NPP) based on energy and water balance.
  • The SSiB was coupled with a dynamic vegetation model, the Top-down Representation of Interactive Foliage and Flora Including Dynamics Model , to calculate leaf area index (LAI), canopy height, and PFT fractional coverage according to the carbon balance (Cox, 2001; Zhang et al., 2015; Harper et al., 2016; Liu et al., 2019).
  • The authors have improved the PFT competition strategy and plant physiology processes to make the SSiB4/TRIFFID suitable for seasonal, interannual, and decadal studies (Zhang et al., 2015; Liu et al., 2019).

2.2 Fire model and modifications

  • A process-based fire model of intermediate complexity has been implemented in the SSiB4/TRIFFID, called SSiB4/TRIFFID-Fire.
  • The fire model is comprised of three parts: fire occurrence, fire spread, and fire impact.

2.2.1 Fire occurrence

  • Nf is the product of the number of potential ignition counts due to both natural causes (In, count s−1 km−2) and human ignitions (Ia, count s−1 km−2), fuel availability (fb), fuel combustibility (fm), and human suppression factors (feo).
  • The authors only consider non-crop fire by excluding the cropland fraction from burning: Nf = (In+ Ia)fbfmfeo(1− )Ag, (2) where Ag is the land area of the grid cell (km2).
  • Fires in the croplands are excluded here due to their small extent within the major fire regions and their relatively low intensity (Bistinas et al., 2014).

2.2.2 Average spread area after fire ignition

  • The average spread area of a fire is assumed to be elliptical in shape, with the ignition point located at one of the foci and the fastest spread occurring along the major axis.
  • The human suppression factor, Fse, reflects the human impact on fire spread through firefighting activities and is parameterized following Li et al. (2013).

2.2.3 Carbon emissions, post-fire mortality, and emissions of aerosols and trace gases

  • As the carbon cycle in current SSiB4/TRIFFID does not explicitly represent the litter carbon storage and decomposition, the authors assume the litter carbon and woody debris account for about 25 % of aboveground biomass for global forest and about 30 % for savanna and grassland based on previous studies (Pan et al., 2011; de Oliveira et al., 2019).
  • CCj is the corresponding combustion completeness for leaf, stem, root, and litter of the j th PFT (Table S2).
  • Meanwhile, fire-induced mortality transfers carbon from uncombusted leaf, stem, and root to litter: ψj = AbCj · (1−CCj ) ·Mj , (14) where Mj = (Mleaf,Mstem,Mroot)j is the corresponding mortality factor for the j th PFT(Table S2).
  • The emissions of trace gases and aerosols can be applied in the atmospheric chemistry model to calculate the production of secondary aerosols, transport of pollutants, and the resultant aerosol direct and indirect effects on climate.

2.2.4 Including the fire effect on the carbon pool

  • When the Li et al. (2012) fire model is coupled with CLM, the vegetation distribution is prescribed using satellite-based Geosci.
  • In SSiB4/TRIFFID-Fire, the fire-related carbon loss due to combustion and post-fire mortality is transferred to changes of PFT fraction based on carbon balance.
  • Cvj is the carbon density (g C km−2); cij is the competition coefficient between the ith and j th PFTs; and γj (s−1) is the constant disturbance representing the loss of PFT fraction due to fires, pests, windthrow, and many other processes.
  • The fire-caused PFT fraction loss results in bare soil for vegetation spreading decided by the competition strategy in TRIFFID.

2.3 Implementing the fire model in SSiB4/TRIFFID

  • In SSiB4/TRIFFID, SSiB4 provides GPP, autotrophic respiration, and other physical variables such as canopy temperature and soil moisture every 3 h for TRIFFID (Fig. 1).
  • TRIFFID accumulates the 3-hourly GPP and respiration and provides biotic carbon, PFT fractional coverage, vegetation height, and LAI every 10 d, which are used to update surface properties (albedo, canopy height, roughness length, and aerodynamic/canopy resistances) in SSiB4.
  • When the fire model is included, it uses the meteorological forcings and physical variables provided by SSiB4 every 3 h and the biophysical properties (PFT fraction and biotic carbon) provided by TRIFFID every 10 d. TRIFFID updates the vegetation dynamics based on carbon balance on day 10, using the net primary production, fireinduced carbon loss, and PFT competition strategy.
  • The updated vegetation dynamics are transferred to SSiB4 to reflect fire effects on surface properties.

3.1 Experimental design

  • Two sets of offline experiments have been conducted using SSiB4/TRIFFID-Fire, which consist of FIRE-ON (SSiB4/TRIFFID-Fire with fire model switched on) and FIRE-OFF (SSiB4/TRIFFID-Fire with fire model switched off).
  • To obtain the initial conditions for these two experiments, similar to their previous SSiB4/TRIFFID experiments (Zhang et al., 2015; Liu et al., 2019), the authors conducted spinup simulations (SPFIRE-ON and SPFIRE-OFF) for 100 years to reach a quasi-equilibrium PFT distribution with and without fire disturbance.
  • These spin-up simulations were initialized using the quasi-equilibrium state from Liu et al. (2019) (SPINIT in Fig. 2) and were driven by climatology forcing averaged over 1948–1972 and atmospheric CO2 concentration, population density, and GDP in 1948 (Fig. 2).
  • FIRE-ON produced the fire regime, ecosystem, and surface conditions during 1948–2014.
  • A FIRE-OFF run, based on SPFIRE-OFF, was carried out with the fire model switched off during 1948–2014.

3.2 Model input and validation data

  • The meteorological forcings used to drive SSiB4/TRIFFIDFire for the period of 1948–2014 are from the Princeton global meteorological dataset for land surface modeling (Sheffield et al., 2006), including surface air temperature, surface pressure, specific humidity, wind speed, downward shortwave radiation flux, downward longwave radiation flux, and precipitation (Table 3).
  • The dataset is constructed by combining global observation-based datasets with the NCEP/NCAR reanalysis.
  • The 2-hourly climatology lightning flashes data from NASA LIS/OTD v2.2 at 2.5◦× 2.5◦ resolution are used to calculate the number of natural ignitions.
  • The GFED fire product provides the burned area and fire emissions on the global scale and has been widely used for fire model validation and calibration (van Marle et al., 2017; Li et al., 2019).

4 Results

  • This section evaluates the model simulation of burned area, carbon emissions, PFT fraction, and GPP by comparing FIRE-ON results with GFED4s, GLC2000, and FLUXNET-MTE data.
  • Specifically, the authors will focus on the model description of fire seasonality.
  • -Fire is applied to assess the long-term fire effect on the ecosystem and surface energy budget using the differences between the FIRE-ON and FIRE-OFF.

4.1 Burned area

  • The simulations of burned area are evaluated using satellitebased product GFED4s for the period of 2000–2014.
  • In 8 out of the 14 sub-regions, SSiB4/TRIFFID-Fire reproduces the observed interannual variability (IAV) of burned area well, with the correlation between simulations and observations significant at p < 0.05 (Fig. 4c).
  • The simulated IAV of SHAF burned area is not as good as other savanna fire regions (e.g., NHAF, SHSA, and AUST), although the IAV is small there (Fig. 4c).
  • As lightning flash is the predominant ignition source in the Northern Hemisphere high latitudes, the application of climatology lightning has a greater impact in boreal regions than in other parts of the globe.
  • -Fire generates two fire seasons in June–July–August and December–January–February, capturing the peak month in August but underestimating the burned area in December.

4.2 Fire emissions

  • Biomass burning emissions are determined by burned area, fuel combustion rate per unit area, and emission factors per unit mass of fuel burned (van der Werf et al., 2017).
  • -Fire captures the high carbon emissions in tropical savannas, the intermediate emissions in Northern Hemisphere boreal forests, and the low emissions in hu- https://doi.org/10.5194/gmd-13-6029-2020.
  • The simulated global CO emission is 433.7.
  • In NHAF, SSiB4/TRIFFID-Fire reproduces the large CO emission in DJF, although the model slightly underestimates CO emis- sion in December and overestimates it in February.

4.3 PFT distribution and GPP

  • The simulation of vegetation coverage, which represents model description of biomass allocation and influences the fuel availability and flammability in fire modeling, is evaluated against GLC2000 (Bartholome and Belward, 2005).
  • Compared with the observations (Fig. 10a), SSiB4/TRIFFID-Fire captures the fractional coverage of trees in the Amazon rainforest, tropical Africa, equatorial Asia, Southeast Asia, southeastern North America, and the Northern Hemisphere Boreal regions (Fig. 10c).
  • Shrubs are primarily located in the semiarid regions and the pan-Arctic area and tundra is limited to the pan-Arctic area and Tibetan Plateau (Fig. S3f–g).
  • The highest GPP occurs in the tropical evergreen forest and decreases with latitude in both observations and model.
  • In addition, the correlation of IAV of global GPP is 0.68 (p < 0.05) between SSiB4/TRIFFID-Fire and FLUXNET-MTE, indicating that the model has reasonably captured the terrestrial ecosystem variability during the historical period.

4.4 Fire effects on ecosystem characteristics and surface properties

  • The authors investigate long-term fire effects on ecosystem characteristics, surface properties, and surface energy budget using the differences between FIREON and FIRE-OFF (FIRE-ON minus FIRE-OFF).
  • In Africa, the simulated fire effects on vegetation structure (tree and grass cover, LAI, and vegetation height) peak in the tropical savanna surrounding the forests and gradually decrease towards the deserts (Fig. 11).
  • Therefore, the latent heat can be increased due to enhanced soil evaporation.

5 Conclusions and discussions

  • The authors have implemented a process-based fire model of intermediate complexity into a DGVM, SSiB4/TRIFFID, which is based on the surface water, carbon, and energy balances, as well as the PFT competition.
  • The SSiB4/TRIFFID-Fire is then applied to study the long-term fire effects on ecosystem characteristics and surface energy.
  • As Li et al. (2017) is the only modeling study investigating the long-term fire effects on the land energy budget, their simulation provides another approach that quantifies fire effects using a different land surface model with different approaches in parameterizing some land surface processes and vegetation dynamics.
  • All authors (HH, YX, FL, and YL) have contributed to the analysis methods and to the text.
  • The authors acknowledge the use of the Cheyenne supercomputer (https://doi.org/10.5065/D6RX99HX, Computational and Information Systems Laboratory, 2019), provided by NCAR CISL, for providing HPC resources.

Did you find this useful? Give us your feedback

Figures (13)

Content maybe subject to copyright    Report

Geosci. Model Dev., 13, 6029–6050, 2020
https://doi.org/10.5194/gmd-13-6029-2020
© Author(s) 2020. This work is distributed under
the Creative Commons Attribution 4.0 License.
Modeling long-term fire impact on ecosystem characteristics and
surface energy using a process-based vegetation–fire
model SSiB4/TRIFFID-Fire v1.0
Huilin Huang
1
, Yongkang Xue
1,2
, Fang Li
3
, and Ye Liu
1
1
Department of Geography, University of California, Los Angeles, CA 90095, USA
2
Department of Atmospheric & Oceanic Sciences, University of California, Los Angeles, CA 90095, USA
3
International Center for Climate and Environmental Sciences, Institute of Atmospheric Physics,
Chinese Academy of Sciences, Beijing, China
Correspondence: Yongkang Xue (yxue@geog.ucla.edu)
Received: 26 April 2020 Discussion started: 15 June 2020
Revised: 19 September 2020 Accepted: 16 October 2020 Published: 2 December 2020
Abstract. Fire is one of the primary disturbances to the
distribution and ecological properties of the world’s ma-
jor biomes and can influence the surface fluxes and climate
through vegetation–climate interactions. This study incorpo-
rates a fire model of intermediate complexity to a biophys-
ical model with dynamic vegetation, SSiB4/TRIFFID (The
Simplified Simple Biosphere Model coupled with the Top-
down Representation of Interactive Foliage and Flora Includ-
ing Dynamics Model). This new model, SSiB4/TRIFFID-
Fire, updating fire impact on the terrestrial carbon cycle ev-
ery 10 d, is then used to simulate the burned area during
1948–2014. The simulated global burned area in 2000–2014
is 471.9 Mha yr
1
, close to the estimate of 478.1 Mha yr
1
in Global Fire Emission Database v4s (GFED4s), with a spa-
tial correlation of 0.8. The SSiB4/TRIFFID-Fire reproduces
temporal variations of the burned area at monthly to inter-
annual scales. Specifically, it captures the observed decline
trend in northern African savanna fire and accurately simu-
lates the fire seasonality in most major fire regions. The sim-
ulated fire carbon emission is 2.19 Pg yr
1
, slightly higher
than the GFED4s (2.07 Pg yr
1
).
The SSiB4/TRIFFID-Fire is applied to assess the long-
term fire impact on ecosystem characteristics and surface en-
ergy budget by comparing model runs with and without fire
(FIRE-ON minus FIRE-OFF). The FIRE-ON simulation re-
duces tree cover over 4.5 % of the global land surface, ac-
companied by a decrease in leaf area index and vegetation
height by 0.10 m
2
m
2
and 1.24 m, respectively. The surface
albedo and sensible heat are reduced throughout the year,
while latent heat flux decreases in the fire season but in-
creases in the rainy season. Fire results in an increase in sur-
face temperature over most fire regions.
1 Introduction
Wildfire, whether natural or human-made, is one of the pri-
mary ecosystem disturbances and it plays a major role in the
terrestrial biogeochemical cycles and ecological succession
across spatial and temporal scales (Sousa, 1984; Bowman et
al., 2009). Every year in the dry season, wildfires burn about
400 Mha of land vegetated areas, leaving behind numerous
scars in the landscape (Giglio et al., 2013; Chuvieco et al.,
2016). Fires affect the climate through modification of wa-
ter, energy, and momentum exchange between land and at-
mosphere (Chambers and Chapin, 2002; Bond-Lamberty et
al., 2009) and can interact with monsoons by affecting at-
mospheric circulations (De Sales et al., 2016; Saha et al.,
2016). Fires are also important sources of global carbon,
aerosols, and trace gas emissions. Based on the latest satel-
lite estimates, global fires emit 1.5–4.2 Pg C yr
1
carbon, 7–
8.2 Pg C yr
1
CO
2
, and 1.9–6.0 Tg C yr
1
black carbon to
the atmosphere (Chuvieco et al., 2016; van der Werf et al.,
2017; Li et al., 2019). Fire emissions contribute to increases
in greenhouse gases and cloud condensation nuclei through
geochemistry processes (Scholes et al., 1996), affecting ra-
diative forcing, the hydrology cycle (Ward et al., 2012; Jiang
Published by Copernicus Publications on behalf of the European Geosciences Union.

6030 H. Huang et al.: Modeling long-term fire impact on ecosystem characteristics
et al., 2016; Hamilton et al., 2018), and air quality (van der
Werf et al., 2010; Johnston et al., 2012).
Since the early 2000s, fire models have been developed
within Dynamic Global Vegetation Models (DGVMs) to ex-
plicitly describe the burned area, fire emissions, and fire dis-
turbance on terrestrial ecosystems (Thonicke et al., 2001;
Venevsky et al., 2002; Arora and Boer, 2005; Thonicke et
al., 2010; Li et al., 2012; Pfeiffer et al., 2013; Lasslop et al.,
2014; Yue et al., 2014; Rabin et al., 2018; Burton et al., 2019;
Venevsky et al., 2019). These fire models have various lev-
els of complexity, from simple statistical models (SIMFIRE;
Knorr et al., 2016) to complicated process-based ones such
as SPITFIRE (Thonicke et al., 2010) and MC2 (Bachelet et
al., 2015). With increasing complexity, more fire processes
and fire characteristics are considered in fire models. In gen-
eral, current fire models broadly capture the global amounts
and spatial distribution of burned area and carbon emissions,
as compared to different observations. However, many em-
pirically determined parameters are included in the compli-
cated process-based models, which leads to large uncertain-
ties. There is no model that outperforms other models across
all fire variables (Hantson et al., 2020). Moreover, current fire
models have deficiencies in simulating the peak fire month,
fire season length, and interannual variability, as reported by
the Fire Model Intercomparison Project (FireMIP; Hantson
et al., 2020; Li et al., 2019). Most fire models show a 1–2
month shift in peak burned area and simulate a longer fire
season compared to observations.
Fire models have been used to reconstruct fire history be-
fore the satellite era (Yang et al., 2015; van Marle et al., 2017;
Li et al., 2019). In addition, they are widely used to attribute
historical variability of burned area to various climate and
anthropogenic driving factors (Kloster et al., 2012; Andela
et al., 2017; Forkel et al., 2019; Teckentrup et al., 2019).
Some fire models have been used to assess long-term fire im-
pact on the terrestrial carbon cycle by comparing a reference
simulation with fire and a sensitivity simulation representing
“a world without fire”. However, the simulated responses of
vegetation and carbon cycle are divergent. Bond et al. (2005)
reported that forest cover would double in a world without
fire, while in the recent fire-coupled DGVMs, a much smaller
tree cover reduction by 10 % (ranges between 3 % and 25 %)
is simulated when fire is taken into account (Lasslop et al.,
2020). Earlier model-based studies reported that fire reduced
terrestrial carbon uptake. However, the range of the quan-
tified reduction was fairly broad (0.05–3.60 Pg C yr
1
), and
most studies did not consider the fire effects on vegetation
distribution and related mechanisms (Li et al., 2014; Yue et
al., 2015; Poulter et al., 2015; Yang et al., 2015; Seo and Kim,
2019; Zou et al. 2020).
Thus far, only the fire model developed by Li et al. (2012,
2013) has been used to investigate the long-term fire effects
on surface energy. By comparing the simulated climate with
and without fire, Li et al. (2017) concluded that fire caused
a significant decrease in surface radiation, latent heat, and
a slight decrease in sensible heat fluxes through changes in
biophysical properties such as albedo, Bowen ratio, and aero-
dynamic resistance. An increase in surface temperature was
found over most fire regions. However, the long-term fire im-
pact on vegetation distribution was not taken into account in
Li et al. (2017), which has been widely observed on site-level
studies (Higgins et al., 2007; Smit et al., 2010) and can cause
substantial changes in aerodynamic resistance due to conver-
sions of dominant plant functional type (PFT) (Huang et al.,
2020b). Moreover, Li et al. (2017) focused on the annual fire
impact on energy fluxes. However, fire’s effects on energy
budget can have large seasonal variations associated with the
vegetation loss during the fire seasons and vegetation recov-
ery during post-fire rainy seasons. The seasonal variations in
fire effects have not been investigated in any fire studies.
In the original SSiB4/TRIFFID, the carbon disturbance
caused by fire and insects was assumed to be a constant,
which depended solely on PFT without spatial and tempo-
ral changes (Cox et al., 2001; Liu et al., 2019). However,
the fire disturbance is varies greatly with climate, vegeta-
tion productivity, and socioeconomic conditions, which has
a strong influence on vegetation dynamics, carbon cycling,
and soil processes. In this study, we develop the fire mod-
eling by incorporating the fire scheme of Li et al. (2012,
2013) to SSiB4/TRIFFID (hereafter, SSiB4/TRIFFID-Fire).
The SSiB4/TRIFFID-Fire model updates fire-induced carbon
loss every 10 d, which has been rarely employed in current
process-based fire models, and is used to provide a quanti-
tative assessment of fire impact on ecosystem characteristics
and surface energy at subseasonal, seasonal, interannual, and
long-term scales. Specifically, our objectives are (1) to evalu-
ate the climatology and interannual variability of burned area
and carbon emissions simulated by offline SSiB4/TRIFFID-
Fire, (2) to assess the ability of SSiB4/TRIFFID-Fire in cap-
turing the fire seasonality in major fire regions, and (3) to
assess the long-term fire impact on PFT distribution and veg-
etation properties and the resultant changes in seasonal sur-
face energy budget and temperature. In Sect. 2, we provide
a brief description of the DGVM, SSiB4/TRIFFID; the fire
model, taken from Li et al. (2012, 2013); and the coupling
procedures. The experimental design and data for model in-
put and validation are introduced in Sect. 3. The fire model
evaluation on a global scale and the application of long-term
fire impact on ecosystem characteristics and surface proper-
ties are presented in Sect. 4. Discussions and conclusions are
given in Sect. 5.
2 Method
2.1 Land and vegetation model
The Simplified Simple Biosphere Model (SSiB, Xue et al.,
1991; Zhan et al., 2003) is a biophysical model which sim-
ulates fluxes of surface radiation, momentum, sensible and
Geosci. Model Dev., 13, 6029–6050, 2020 https://doi.org/10.5194/gmd-13-6029-2020

H. Huang et al.: Modeling long-term fire impact on ecosystem characteristics 6031
latent heat, runoff, soil moisture, surface temperature, and
vegetation gross and net primary productivity (GPP and
NPP) based on energy and water balance. The SSiB was
coupled with a dynamic vegetation model, the Top-down
Representation of Interactive Foliage and Flora Including
Dynamics Model (TRIFFID), to calculate leaf area index
(LAI), canopy height, and PFT fractional coverage accord-
ing to the carbon balance (Cox, 2001; Zhang et al., 2015;
Harper et al., 2016; Liu et al., 2019). We have improved
the PFT competition strategy and plant physiology processes
to make the SSiB4/TRIFFID suitable for seasonal, interan-
nual, and decadal studies (Zhang et al., 2015; Liu et al.,
2019). SSiB4/TRIFFID includes seven PFTs: (1) broadleaf
evergreen trees (BET), (2) needleleaf evergreen trees (NET),
(3) broadleaf deciduous trees (BDT), (4) C
3
grasses, (5) C
4
plants, (6) shrubs, and (7) tundra. The coverage of a PFT is
determined by net carbon availability, competition between
species, and disturbance, which implicitly includes mortal-
ity due to fires, pests, and windthrow. A detailed description
and validation of SSiB4/TRIFFID can be found in Zhang et
al. (2015) and Liu et al. (2019).
2.2 Fire model and modifications
In this study, a process-based fire model of intermediate
complexity has been implemented in the SSiB4/TRIFFID,
called SSiB4/TRIFFID-Fire. The fire model developed by
Li et al. (2012, 2013) was first built on the model platform
of CLM-DGVM and has been incorporated in IAP-DGVM
(Zeng et al., 2014), CLM4.5 (Oleson et al., 2013), CLM5
(Lawrence et al., 2019), LM3 in Earth system model GFDL-
ESM (Rabin et al., 2018; Ward et al., 2018), AVIM in Cli-
mate System Model BCC-CSM (Weiping Li, personal com-
munication, 2016), E3SM Land Model (ELM; Ricciuto et al.,
2018), NASA GEOS catchment-CN4.5 model (Zeng et al.,
2019), and DLEM (Yang et al., 2014), and it has been partly
used in GLASS-CTEM (Melton and Arora, 2016). The fol-
lowing briefly describes the fire schemes adapted from Li et
al. (2012, 2013), Li and Lawrence (2017), and our own mod-
ifications.
The fire model is comprised of three parts: fire occurrence,
fire spread, and fire impact. The basic equation is that the
burned area in a grid cell (A
b
, km
2
s
1
) is determined by the
number of fires per time step (N
f
, count s
1
) and the average
spread area per fire (a, km
2
count
1
):
A
b
= N
f
a. (1)
2.2.1 Fire occurrence
N
f
is the product of the number of potential ignition counts
due to both natural causes (I
n
, count s
1
km
2
) and human
ignitions (I
a
, count s
1
km
2
), fuel availability (f
b
), fuel
combustibility (f
m
), and human suppression factors (f
eo
). In
this paper, we only consider non-crop fire by excluding the
cropland fraction (f
crop
) from burning:
N
f
= (I
n
+ I
a
)f
b
f
m
f
eo
(1 f
crop
)A
g
, (2)
where A
g
is the land area of the grid cell (km
2
). Fires in the
croplands are excluded here due to their small extent within
the major fire regions and their relatively low intensity (Bisti-
nas et al., 2014). Cropland fire is still a major uncertainty in
remote sensing datasets (Randerson et al., 2012), and more
data and investigation are needed.
The number of natural ignitions is related to lightning
flashes (I
l
, count s
1
); cloud-to-ground lightning fraction,
1
5.16+2.16 cos[3,min(60,λ)]
, which depends on latitude λ (Pren-
tice and Mackerras, 1977); and ignition efficiency (ψ =
0.22). The anthropogenic ignition, I
a
, is parameterized using
the number of potential anthropogenic ignitions by a person
(α = 1.35 × 10
9
count per person s
1
) and population den-
sity (D
p
; person) (Venevsky et al. 2002):
I
n
=
ψ
5.16 + 2.16cos[3,min(60,λ)]
I
l
, (3)
I
a
= αD
p
× (6.8D
0.6
p
). (4)
The fuel availability f
b
(fraction, range 01) is given as fol-
lows:
f
b
=
0 B
ag
B
low
B
ag
B
low
B
up
B
low
B
low
< B
ag
< B
up
1 B
ag
B
up
, (5)
where B
ag
(g C m
2
) is the aboveground biomass (leaf and
stem in SSiB4/TRIFFID-Fire) of all PFTs. Following Li et
al. (2012), we use B
low
= 155 g C m
2
as the lower fuel
threshold, below which fire does not occur and B
up
=
1050 g C m
2
as the upper fuel threshold, above which fuel
load is not a constraint for fire occurrence.
Fuel combustibility f
m
(fraction, 0–1) is given as follow:
f
m
= f
RH
f
θ
, (6)
where f
RH
and f
θ
represent the dependence of fuel com-
bustibility on relative humidity (RH; %) and the root zone
soil moisture (θ), respectively (Li and Lawrence, 2017). Fol-
lowing Li et al. (2013), we assume f
m
= 0 when surface air
temperature T is below 10
C. f
RH
reflects the impact of
real-time climate conditions on fuel combustibility, while f
θ
reflects the response of fuel combustibility to the preceding
climate conditions (Shinoda and Yamaguchi, 2003):
f
RH
=
0 RH RH
up
(
RH
up
RH
RH
up
RH
low
)
1.3
RH
low
< RH < RH
up
1 RH RH
low
, (7)
f
θ
=
0 θ θ
up
(
θ
up
θ
θ
up
θ
low
)
0.7
θ
low
< θ < θ
up
1 θ θ
low
. (8)
Relative humidity suppresses fire occurrence when it is larger
than RH
up
= 70 %, and relative humidity does not con-
https://doi.org/10.5194/gmd-13-6029-2020 Geosci. Model Dev., 13, 6029–6050, 2020

6032 H. Huang et al.: Modeling long-term fire impact on ecosystem characteristics
Table 1. The upper (θ
up
) and lower (θ
low
) thresholds of root zone
soil moisture for PFTs in SSiB4/TRIFFID-Fire.
Vegetation types θ
up
θ
low
BET 0.80 0.30
NET 0.80 0.30
BDT 0.80 0.30
C
3
grasses 0.75 0.30
C
4
plants 0.75 0.30
Shrubs 0.60 0.30
Tundra 0.60 0.30
strain fire when it is smaller than RH
low
= 30 %. The PFT-
dependent θ
up
and θ
low
are used as the upper and lower
thresholds of soil moisture in a similar way to the thresh-
olds of relative humidity (Table 1). In the Li et al. (2012)
fire scheme, this factor (f
θ
) is parameterized using root zone
soil moisture potential factor β (0–1.0), a model-dependent
variable used to calculate transpiration in CLM (Li and
Lawrence, 2017). β
low
= 0.85 and β
up
= 0.98 are used as
the lower and upper thresholds for all PFTs, yet the narrow
range of β in CLM5 has led to fire model too sensitive to
drought. In SSiB4/TRIFFID-Fire, the root zone soil moisture
θ is found to be the best variable to describe the dependence
of fuel combustibility on the preceding climate.
The human suppression factor (f
eo
; 0–1) reflects the de-
mographic (f
d
) and economic (f
e
) impact on fire occurrence
in populated areas (population density D
p
> 0.1 per person
km
2
):
f
seo
= f
d
f
e
. (9)
The human suppression is assumed to be negligible (f
seo
=
1) when D
p
0.1 person km
2
. A detailed description of
f
d
and f
e
parameterization can be found in Li et al. (2012,
2013).
2.2.2 Average spread area after fire ignition
The average spread area of a fire is assumed to be elliptical
in shape, with the ignition point located at one of the foci
and the fastest spread occurring along the major axis. The
average burned area of a fire a (km
2
per count) is represented
as follows (Li et al., 2012):
a = π L
B
(u
max
g
0
F
m
τ/1000)
2
F
se
, (10)
where L
B
is the length-to-breadth ratio of the ellipse shape
and is related to the wind speed, W (m s
1
):
L
B
= 1.0 + 10.0
1 exp
(
0.06W
)
. (11)
u
max
is the PFT-dependent maximum fire spread rate (m s
1
;
Table S1). g
0
= 0.05 is the dependence of fire spread rate
perpendicular to the wind direction. F
m
is the influence of
fuel wetness on fire spread and is assumed to be related to
f
m
in the fire occurrence Eq. (6):
F
m
= f
0.5
m
. (12)
τ (= 3600 × 24 s) is the global average fire duration, which
is simply taken to be 1 d, as reported by Giglio et al. (2006).
The human suppression factor, F
se
, reflects the human impact
on fire spread through firefighting activities and is parameter-
ized following Li et al. (2013).
2.2.3 Carbon emissions, post-fire mortality, and
emissions of aerosols and trace gases
In post-fire regions, the fire carbon emission, ϕ
j
(g C s
1
),
from the jth PFT is calculated based on the burned area (A
b
;
km
2
s
1
), Carbon density C
j
, and carbon combustion com-
pleteness CC
j
:
ϕ
j
= A
b
C
j
· CC
j
. (13)
C
j
= (C
leaf
, C
stem
, C
root
, C
litter
)
j
is carbon density vector
(g C km
2
) for leaf, stem, root, and litter of the j th PFT
calculated in TRIFFID. As the carbon cycle in current
SSiB4/TRIFFID does not explicitly represent the litter car-
bon storage and decomposition, we assume the litter carbon
and woody debris account for about 25 % of aboveground
biomass for global forest and about 30 % for savanna and
grassland based on previous studies (Pan et al., 2011; de
Oliveira et al., 2019). CC
j
is the corresponding combustion
completeness for leaf, stem, root, and litter of the jth PFT
(Table S2). Meanwhile, fire-induced mortality transfers car-
bon from uncombusted leaf, stem, and root to litter:
ψ
j
= A
b
C
j
· (1 CC
j
) · M
j
, (14)
where M
j
= (M
leaf
, M
stem
, M
root
)
j
is the corresponding
mortality factor for the jth PFT(Table S2).
Finally, the emissions of trace gases and aerosols species
x for the j th PFT (EM
x,j
, g s
1
) can be calculated from car-
bon emissions (ϕ
j
) using the PFT-dependent emission factor
(EF
x,j
, g species (kg dm)
1
):
EM
x,j
= EF
x,j
ϕ
j
[C]
, (15)
where
[
C
]
(= 0.5 g C (kg dm)
1
) is a unit conversion factor
from dry matter to carbon (Li et al., 2019). The emission
factors, EF
x,j
, of trace gases and aerosols in Table S3 are
based on Andreae (2019). The emissions of trace gases and
aerosols can be applied in the atmospheric chemistry model
to calculate the production of secondary aerosols, transport
of pollutants, and the resultant aerosol direct and indirect ef-
fects on climate.
2.2.4 Including the fire effect on the carbon pool
When the Li et al. (2012) fire model is coupled with CLM,
the vegetation distribution is prescribed using satellite-based
Geosci. Model Dev., 13, 6029–6050, 2020 https://doi.org/10.5194/gmd-13-6029-2020

H. Huang et al.: Modeling long-term fire impact on ecosystem characteristics 6033
land cover, and therefore the fire impact on vegetation cover
is not simulated. In SSiB4/TRIFFID-Fire, the fire-related
carbon loss due to combustion and post-fire mortality is
transferred to changes of PFT fraction based on carbon bal-
ance.
In TRIFFID (Cox, 2001), the fractional change of the jth
PFT (
df
j
dt
) is governed by the Lotka–Volterra equation:
df
j
dt
=
λ
j
NPP
j
f
j
Cv
j
h
1
X
j
c
ij
f
j
i
γ
j
f
j
, (16)
where f
j
is the fractional coverage of the j th PFT; λ
j
NPP
j
is
the carbon available for spreading; Cv
j
is the carbon density
(g C km
2
); c
ij
is the competition coefficient between the ith
and jth PFTs; and γ
j
(s
1
) is the constant disturbance repre-
senting the loss of PFT fraction due to fires, pests, windthrow,
and many other processes.
When the fire model is coupled to SSiB4/TRIFFID, the
loss of PFT fraction due to fires (β
j
) can be explicitly derived
from the fire-induced carbon loss:
β
j
=
j
+ ψ
j
) · f
j
Cv
j
, (17)
where ϕ
j
and ψ
j
are PFT-dependent carbon loss due to com-
bustion and post-fire mortality, respectively. The fire-caused
PFT fraction loss results in bare soil for vegetation spreading
decided by the competition strategy in TRIFFID. As such,
fire disturbance is explicitly represented and varies in space
and time, and the original γ
j
is adjusted to γ
0
j
to exclude fire
disturbance (Table 2):
df
j
dt
=
λ
j
· NPP
j
· f
j
C
j
h
1
X
j
c
ij
f
j
i
γ
0
j
+ β
i
f
j
. (18)
2.3 Implementing the fire model in SSiB4/TRIFFID
In SSiB4/TRIFFID, SSiB4 provides GPP, autotrophic res-
piration, and other physical variables such as canopy tem-
perature and soil moisture every 3 h for TRIFFID (Fig. 1).
TRIFFID accumulates the 3-hourly GPP and respiration and
provides biotic carbon, PFT fractional coverage, vegetation
height, and LAI every 10 d, which are used to update sur-
face properties (albedo, canopy height, roughness length, and
aerodynamic/canopy resistances) in SSiB4. When the fire
model is included, it uses the meteorological forcings and
physical variables provided by SSiB4 every 3 h and the bio-
physical properties (PFT fraction and biotic carbon) provided
by TRIFFID every 10 d. The fire model calculates the burned
area, carbon combustion, post-fire mortality, and emissions
every 3 h, and the fire-induced carbon loss is subtracted from
fuel load. The carbon loss is accumulated within 10 d in
the fire model and is transferred to TRIFFID on day 10.
TRIFFID updates the vegetation dynamics based on carbon
balance on day 10, using the net primary production, fire-
induced carbon loss, and PFT competition strategy. The up-
dated vegetation dynamics are transferred to SSiB4 to reflect
fire effects on surface properties.
3 Experimental setup and data
3.1 Experimental design
Two sets of offline experiments have been conducted
using SSiB4/TRIFFID-Fire, which consist of FIRE-ON
(SSiB4/TRIFFID-Fire with fire model switched on) and
FIRE-OFF (SSiB4/TRIFFID-Fire with fire model switched
off). To obtain the initial conditions for these two experi-
ments, similar to our previous SSiB4/TRIFFID experiments
(Zhang et al., 2015; Liu et al., 2019), we conducted spin-
up simulations (SP
FIRE-ON
and SP
FIRE-OFF
) for 100 years to
reach a quasi-equilibrium PFT distribution with and with-
out fire disturbance. These spin-up simulations were initial-
ized using the quasi-equilibrium state from Liu et al. (2019)
(SP
INIT
in Fig. 2) and were driven by climatology forcing av-
eraged over 1948–1972 and atmospheric CO
2
concentration,
population density, and GDP in 1948 (Fig. 2). Following Liu
et al. (2019), the quasi-equilibrium status is defined as the
rate of relative change in fractional coverage of all PFTs is
less than 2 % over the last 10 years of simulation.
Based on the quasi-equilibrium status with fire disturbance
(SP
FIRE-ON
), a transient run was performed (FIRE-ON) with
the fire model turned on from 1948 to 2014 (Fig. 2). The
model was forced by 3-hourly meteorological forcings and
yearly updated atmospheric CO
2
concentration, population
density, and GDP data. FIRE-ON produced the fire regime,
ecosystem, and surface conditions during 1948–2014. A
FIRE-OFF run, based on SP
FIRE-OFF
, was carried out with
the fire model switched off during 1948–2014. The vegeta-
tion distribution was allowed to respond to climate variations
in both FIRE-ON and FIRE-OFF simulations and to fire dis-
turbances only in the FIRE-ON.
3.2 Model input and validation data
The meteorological forcings used to drive SSiB4/TRIFFID-
Fire for the period of 1948–2014 are from the Prince-
ton global meteorological dataset for land surface model-
ing (Sheffield et al., 2006), including surface air tempera-
ture, surface pressure, specific humidity, wind speed, down-
ward shortwave radiation flux, downward longwave radia-
tion flux, and precipitation (Table 3). The dataset is con-
structed by combining global observation-based datasets
with the NCEP/NCAR reanalysis. The spatial resolution is
1.0
× 1.0
, and the temporal interval is 3 h.
The required inputs for driving the fire model are listed
in Table 3. The 2-hourly climatology lightning flashes data
from NASA LIS/OTD v2.2 at 2.5
× 2.5
resolution are used
to calculate the number of natural ignitions. The population
https://doi.org/10.5194/gmd-13-6029-2020 Geosci. Model Dev., 13, 6029–6050, 2020

Citations
More filters
01 Dec 2012
Abstract: We upscaled FLUXNET observations of carbon dioxide, water, and energy fluxes to the global scale using the machine learning technique, model tree ensembles (MTE). We trained MTE to predict site-level gross primary productivity (GPP), terrestrial ecosystem respiration (TER), net ecosystem exchange (NEE), latent energy (LE), and sensible heat (H) based on remote sensing indices, climate and meteorological data, and information on land use. We applied the trained MTEs to generate global flux fields at a 0.5 degrees x 0.5 degrees spatial resolution and a monthly temporal resolution from 1982 to 2008. Cross-validation analyses revealed good performance of MTE in predicting among-site flux variability with modeling efficiencies (MEf) between 0.64 and 0.84, except for NEE (MEf = 0.32). Performance was also good for predicting seasonal patterns (MEf between 0.84 and 0.89, except for NEE (0.64)). By comparison, predictions of monthly anomalies were not as strong (MEf between 0.29 and 0.52). Improved accounting of disturbance and lagged environmental effects, along with improved characterization of errors in the training data set, would contribute most to further reducing uncertainties. Our global estimates of LE (158 +/- 7 J x 10(18) yr(-1)), H (164 +/- 15 J x 10(18) yr(-1)), and GPP (119 +/- 6 Pg C yr(-1)) were similar to independent estimates. Our global TER estimate (96 +/- 6 Pg C yr(-1)) was likely underestimated by 5-10%. Hot spot regions of interannual variability in carbon fluxes occurred in semiarid to semihumid regions and were controlled by moisture supply. Overall, GPP was more important to interannual variability in NEE than TER. Our empirically derived fluxes may be used for calibration and evaluation of land surface process models and for exploratory and diagnostic assessments of the biosphere.

948 citations

Journal ArticleDOI
TL;DR: In this paper, the authors provide global estimates of these bulk aerodynamic parameters, including d, z0, rb, and rd, for the period 1982-2017 based on remote-sensed leaf area index (LAI), h, and plant functional type dependent canopy morphological characteristics.

21 citations

15 Dec 2014
TL;DR: In this article, the authors investigated the impact of burned areas on the surface energy balance and monthly precipitation in northern Africa as simulated by a state-of-the-art regional model and found that post-fire land condition deterioration resulted in a decrease in precipitation over sub-Saharan Africa, associated with the weakening of the West African monsoon progression through the region.
Abstract: This study investigates the impact of burned areas on the surface energy balance and monthly precipitation in northern Africa as simulated by a state-of-the-art regional model. Mean burned area fraction derived from MODIS date of burning product was implemented in a set of 1-year long WRF-NMM/SSiB2 model simulations. Vegetation cover fraction and LAI were degraded daily based on mean burned area fraction and on the survival rate for each vegetation land cover type. Additionally, ground darkening associated with wildfire-induced ash and charcoal deposition was imposed through lower ground albedo for a period after burning. In general, wildfire-induced vegetation and ground condition deterioration increased mean surface albedo by exposing the brighter bare ground, which in turn caused a decrease in monthly surface net radiation. On average, the wildfire-season albedo increase was approximately 6.3 % over the Sahel. The associated decrease in surface available energy caused a drop in surface sensible heat flux to the atmosphere during the dry months of winter and early spring, which gradually transitioned to a more substantial decrease in surface evapotranspiration in April and May that lessened throughout the rainy season. Overall, post-fire land condition deterioration resulted in a decrease in precipitation over sub-Saharan Africa, associated with the weakening of the West African monsoon progression through the region. A decrease in atmospheric moisture flux convergence was observed in the burned area simulations, which played a dominant role in reducing precipitation in the area, especially in the months preceding the monsoon onset. The areas with the largest precipitation impact were those covered by savannas and rainforests, where annual precipitation decreased by 3.8 and 3.3 %, respectively. The resulting precipitation decrease and vegetation deterioration caused a drop in gross primary productivity in the region, which was strongest in late winter and early spring. This study suggests the cooling and drying of atmosphere induced by burned areas caused the strengthening of subsidence during pre-onset and weakening of upward atmospheric motion during onset and mature stages of the monsoon leading to a waning of convective instability and precipitation. Monthly mid-tropospheric vertical wind showed a strengthening of downward motion in winter and spring seasons, and weakening of upward movement during the rainy months. Furthermore, precipitation energy analysis revealed that most of precipitation decrease originated from convective events, which supports the hypothesis of reduced convective instability due to wildfires.

17 citations

Journal ArticleDOI
20 May 2021-PLOS ONE
TL;DR: In this article, the authors identify the potential extent of the tropical dry forest biome based on bioclimatic definitions and climatic data sets to improve global estimates of distribution, cover, and change.
Abstract: There is a debate concerning the definition and extent of tropical dry forest biome and vegetation type at a global spatial scale. We identify the potential extent of the tropical dry forest biome based on bioclimatic definitions and climatic data sets to improve global estimates of distribution, cover, and change. We compared four bioclimatic definitions of the tropical dry forest biome–Murphy and Lugo, Food and Agriculture Organization (FAO), DryFlor, aridity index–using two climatic data sets: WorldClim and Climatologies at High-resolution for the Earth’s Land Surface Areas (CHELSA). We then compared each of the eight unique combinations of bioclimatic definitions and climatic data sets using 540 field plots identified as tropical dry forest from a literature search and evaluated the accuracy of World Wildlife Fund tropical and subtropical dry broadleaf forest ecoregions. We used the definition and climate data that most closely matched field data to calculate forest cover in 2000 and change from 2001 to 2020. Globally, there was low agreement (< 58%) between bioclimatic definitions and WWF ecoregions and only 40% of field plots fell within these ecoregions. FAO using CHELSA had the highest agreement with field plots (81%) and was not correlated with the biome extent. Using the FAO definition with CHELSA climatic data set, we estimate 4,931,414 km 2 of closed canopy (≥ 40% forest cover) tropical dry forest in 2000 and 4,369,695 km 2 in 2020 with a gross loss of 561,719 km 2 (11.4%) from 2001 to 2020. Tropical dry forest biome extent varies significantly based on bioclimatic definition used, with nearly half of all tropical dry forest vegetation missed when using ecoregion boundaries alone, especially in Africa. Using site-specific field validation, we find that the FAO definition using CHELSA provides an accurate, standard, and repeatable way to assess tropical dry forest cover and change at a global scale.

12 citations

Journal ArticleDOI
TL;DR: The proposed BASNet significantly outperforms the state-of-the-art SOD methods quantitatively and qualitatively and achieves a promising prediction speed for processing high-resolution UAV images, thus providing wide-ranging applicability in post-disaster monitoring and management.
Abstract: Since remote sensing images of post-fire vegetation are characterized by high resolution, multiple interferences, and high similarities between the background and the target area, it is difficult for existing methods to detect and segment the burned area in these images with sufficient speed and accuracy. In this article, we apply salient object detection (SOD) to burned area segmentation (BAS), the first time this has been done, and propose an efficient burned area segmentation network (BASNet) to improve the performance of unmanned aerial vehicle (UAV) high-resolution image segmentation. BASNet comprises positioning module and refinement module. The positioning module efficiently extracts high-level semantic features and general contextual information via global average pooling layer and convolutional block (CB) to determine the coarse location of the salient region. The refinement module adopts the CB attention module to effectively discriminate the spatial location of objects. In addition, to effectively combine edge information with spatial location information in the lower layer of the network and the high-level semantic information in the deeper layer, we design the residual fusion module to perform feature fusion by level to obtain the prediction results of the network. Extensive experiments on two UAV datasets collected from Chongli in China and Andong in South Korea, demonstrate that our proposed BASNet significantly outperforms the state-of-the-art SOD methods quantitatively and qualitatively. BASNet also achieves a promising prediction speed for processing high-resolution UAV images, thus providing wide-ranging applicability in post-disaster monitoring and management.

6 citations

References
More filters
01 Dec 2018
TL;DR: In this paper, the authors used a biophysical dynamic vegetation model to estimate large-scale trends in terms of carbon fixation, vegetation growth, and expansion during the period of 1958-2007, and attribute these changes to environmental drivers including elevated atmospheric CO2 concentration, global warming, and climate variability.
Abstract: . The climate regime shift during the 1980s had a substantial impact on the terrestrial ecosystems and vegetation at different scales. However, the mechanisms driving vegetation changes, before and after the shift, remain unclear. In this study, we used a biophysical dynamic vegetation model to estimate large-scale trends in terms of carbon fixation, vegetation growth, and expansion during the period 1958–2007, and to attribute these changes to environmental drivers including elevated atmospheric CO2 concentration (hereafter eCO2 ), global warming, and climate variability (hereafter CV). Simulated leaf area index (LAI) and gross primary production (GPP) were evaluated against observation-based data. Significant spatial correlations are found (correlations > 0.87), along with regionally varying temporal correlations of 0.34–0.80 for LAI and 0.45–0.83 for GPP. More than 40 % of the global land area shows significant positive (increase) or negative (decrease) trends in LAI and GPP during 1958–2007. Regions over the globe show different characteristics in terms of ecosystem trends before and after the 1980s. While 11.7 % and 19.3 % of land have had consistently positive LAI and GPP trends, respectively, since 1958, 17.1 % and 20.1 % of land saw LAI and GPP trends, respectively, reverse during the 1980s. Vegetation fraction cover (FRAC) trends, representing vegetation expansion and/or shrinking, are found at the edges of semi-arid areas and polar areas. Environmental drivers affect the change in ecosystem trend over different regions. Overall, eCO2 consistently contributes to positive LAI and GPP trends in the tropics. Global warming mostly affects LAI, with positive effects in high latitudes and negative effects in subtropical semi-arid areas. CV is found to dominate the variability of FRAC, LAI, and GPP in the semi-humid and semi-arid areas. The eCO2 and global warming effects increased after the 1980s, while the CV effect reversed during the 1980s. In addition, plant competition is shown to have played an important role in determining which driver dominated the regional trends. This paper presents new insight into ecosystem variability and changes in the varying climate since the 1950s.

12 citations

Journal ArticleDOI
TL;DR: In this paper, the authors investigated the effect of interactive processes of fire and ecological succession on surface carbon and water fluxes and vegetation distribution, and found that fire-induced changes in vegetation modify the soil moisture profile because grasslands are more dominant in post fire regions.
Abstract: . Fire plays an important role in terrestrial ecosystems. The burning of biomass affects carbon and water fluxes and vegetation distribution. To understand the effect of interactive processes of fire and ecological succession on surface carbon and water fluxes, this study employed the Community Land Model version 4.5 to conduct a series of experiments that included and excluded fire and dynamic vegetation processes. Results of the experiments that excluded the vegetation dynamics showed a global increase in net ecosystem production (NEP) in post-fire regions, whereas the inclusion of vegetation dynamics revealed a fire-induced decrease in NEP in some regions, which was depicted when the dominant vegetation type was changed from trees to grass. Carbon emissions from fires are enhanced by reduction in NEP when vegetation dynamics are considered; however, this effect is somewhat mitigated by the increase in NEP when vegetation dynamics are not considered. Fire-induced changes in vegetation modify the soil moisture profile because grasslands are more dominant in post-fire regions. This results in less moisture within the top soil layer than that in unburned regions, even though transpiration is reduced overall. These findings are different from those of previous fire model evaluations that ignored vegetation dynamics and thus highlight the importance of interactive processes between fires and vegetation dynamics in evaluating recent model developments.

12 citations

Journal ArticleDOI
TL;DR: The distribution of stocks in the different compartments above the ground points to the fragility of this environment to the random fire effect, common in the region, and shows the need for conservation measures for vegetation maintenance and soil protection to preserve adequate nutrient cycling in the ecosystem.
Abstract: The Cerrado is the second largest biome in Brazil and the most biodiverse tropical savannah in the world and acts as a great sequester of atmospheric carbon. The lack of studies related to the quantification of its total biomass compromises the understanding of the dynamics of CO2 in this biome. Thus, it is relevant to develop studies aiming at obtaining accurate estimates of the carbon stock in the different phytophysiognomies that make the Cerrado, to include them in appropriate forest management models. Based on the hypothesis that the amount of carbon stored can vary according to the vegetation typology and vegetation compartments, the aerial stock of dry biomass and carbon were estimated in different compartments (arboreal, herbaceous-shrub and litter). The study was developed in open Brazilian savannah and soils on the sandstone and discussed the effect of fire on this phytophysiognomy. For the arboreal compartment were adjusted mathematical models to fit the biomass equations to estimate the individual stock of the trees in this compartment. The results of the stocks were discussed considering the effect of fire on the phytophysiognomy. Based on the precision and extra distribution measures, the Schumacher-Hall (non-logarithmic) equation presented better results to estimate the individual biomass and carbon stocks of the open Brazilian savannah trees. The aboveground biomass was 12.88 Mg ha−1, corresponding to a total carbon stock of 5.91 Mg ha−1, where most of the stocks are in the herbaceous-shrub compartment (44%). The arboreal compartment accounts for the smallest part of the stocks, followed by the litter. The observed values are in the interval verified for other areas of savannah studied in Brazil. The values verified for the open Brazilian savannah in sandy soils are at the lower limit of this range, due to the nutrient-poor nature of this type of soil. The distribution of stocks in the different compartments above the ground points to the fragility of this environment to the random fire effect, common in the region. That shows the need for conservation measures for vegetation maintenance and soil protection to preserve adequate nutrient cycling in the ecosystem.

6 citations

Frequently Asked Questions (17)
Q1. What have the authors contributed in "Modeling long-term fire impact on ecosystem characteristics and surface energy using a process-based vegetation–fire model ssib4/triffid-fire v1.0" ?

This study incorporates a fire model of intermediate complexity to a biophysical model with dynamic vegetation, SSiB4/TRIFFID ( The Simplified Simple Biosphere Model coupled with the Topdown Representation of Interactive Foliage and Flora Including Dynamics Model ). 

It reasonably reproduces the global GPP and PFT distribution, which is important to study fire effects on the ecosystem. The SSiB4/TRIFFID-Fire is then applied to study the long-term fire effects on ecosystem characteristics and surface energy. HH conducted the simulation with suggestions from FL and YL. 

Other sources of uncertainties include the differences in the partitioning between latent heat and sensible heat fluxes in land surface models, the differences in the parameterization of the evaporation processes, and the changes due to atmospheric feedbacks, such as cloud cover and precipitation changes. 

Over the African and South American savanna, the authors find fire has reduced the area-averaged LAI and vegetation height by 0.52 m2 m−2 (12.5 %) and 5.76 m (49.1 %), respectively. 

Some fire models have been used to assess long-term fire impact on the terrestrial carbon cycle by comparing a reference simulation with fire and a sensitivity simulation representing “a world without fire”. 

In TRIFFID (Cox, 2001), the fractional change of the j th PFT ( dfjdt ) is governed by the Lotka–Volterra equation:dfj dt = λjNPPj fj Cvj[ 1− ∑ j cijfj ] − γjfj , (16)where fj is the fractional coverage of the j th PFT; λjNPPj is the carbon available for spreading; 

The inaccurate simulation of fire season in several fire regions could come from deficiency of the forcing data, the inaccuracy in dynamic vegetation processes, or some processes that control the fire but are not represented in the model. 

When the fire model is coupled to SSiB4/TRIFFID, the loss of PFT fraction due to fires (βj ) can be explicitly derived from the fire-induced carbon loss:βj = (ϕj +ψj ) · fjCvj , (17)where ϕj and ψj are PFT-dependent carbon loss due to combustion and post-fire mortality, respectively. 

The average spread area of a fire is assumed to be elliptical in shape, with the ignition point located at one of the foci and the fastest spread occurring along the major axis. 

In the Li et al. (2012) fire scheme, this factor (fθ ) is parameterized using root zone soil moisture potential factor β (0–1.0), a model-dependent variable used to calculate transpiration in CLM (Li and Lawrence, 2017). 

Cvj is the carbon density (g C km−2); cij is the competition coefficient between the ith and j th PFTs; and γj (s−1) is the constant disturbance representing the loss of PFT fraction due to fires, pests, windthrow, and many other processes. 

(13)Cj = (Cleaf,Cstem,Croot,Clitter)j is carbon density vector (g C km−2) for leaf, stem, root, and litter of the j th PFT calculated in TRIFFID. 

The burned area in African savanna accounts for more than 60 % of the global burned area in both GFED4s and SSiB4/TRIFFID-Fire (Fig. 4b). 

Fire models have been used to reconstruct fire history before the satellite era (Yang et al., 2015; van Marle et al., 2017; Li et al., 2019). 

In SSiB4/TRIFFID-Fire, the root zone soil moisture θ is found to be the best variable to describe the dependence of fuel combustibility on the preceding climate. 

Their results are consistent with the long-term fire experiments that reported that fire strongly affected vegetation structure, lowering the proportions of trees to fire-resistant grasses and reducing the vegetation height and aboveground biomass (Shackleton and Scholes, 2000; Higgins et al., 2007; van Wilgen et al., 2007; Furley et al., 2008; Smit et al., 2010; Devine et al., 2015), and that fire impact is more significant in wetter savanna than in drier savanna (Moreira, 2000; Sankaran et al., 2005). 

In general, the spatial distribution of carbon emissions coincides with that of the burned area: SHAF, NHAF, and SHSA are the major fire emission regions and they contribute to 65.4 % of the total emission in both GFED4s and SSiB4/TRIFFID-Fire (Fig. 8a).