ArticlePDF Available

Atmospheric dispersion and ground deposition induced by the Fukushima Nuclear Power Plant accident: A local-scale simulation and sensitivity study

Authors:

Abstract and Figures

Following the Fukushima Daiichi Nuclear Power Plant (FNPP1) accident on March 2011, radioactive products were released in the atmosphere. Simulations at local scale (within 80 km of FNPP1) were carried out by the Institute of Radiation Protection and Nuclear Safety (IRSN) with the Gaussian Puff model pX, during the crisis and since then, to assess the radiological and environmental consequences. The evolution of atmospheric and ground activity simulated at local scale is presented with a “reference” simulation, whose performance is assessed through comparisons with environmental monitoring data (gamma dose rate and deposition). The results are within a factor of 2–5 of the observations for gamma dose rates (0.52 and 0.85 for FAC2 and FAC5), and 5–10 for deposition (0.31 for FAC2, 0.73 for FAC5 and 0.90 for FAC10). A sensitivity analysis is also made to highlight the most sensitive parameters. A source term comparison is made between IRSN's estimation, and those from Katata et al. (2012) and Stohl et al. (2011). Results are quite sensitive to the source term, but also to wind direction and dispersion parameters. Dry deposition budget is more sensitive than wet deposition. Gamma dose rates are more sensitive than deposition, in particular peak values.
Content may be subject to copyright.
Atmospheric dispersion and ground deposition induced by the
Fukushima Nuclear Power Plant accident: A local-scale simulation
and sensitivity study
I. Korsakissok
*
, A. Mathieu, D. Didier
Institut de Radioprotection et de Sûreté Nucléaire (IRSN), PRP-CRI, SESUC, BMTA, Fontenay-aux-Roses 92262, France
highlights
<We model atmospheric dispersion of the Fukushima accident at local scale (80 km).
<Comparisons to gamma dose rate measurements are within a factor 2e5.
<Comparisons to deposition measurements are within a factor 5e10.
<Source term is the most sensitive parameter.
<Gamma dose rate, especially peak values, are more sensitive than deposition.
article info
Article history:
Received 15 October 2012
Received in revised form
21 December 2012
Accepted 3 January 2013
Keywords:
Atmospheric dispersion
Radionuclides
Fukushima
Model evaluation
Gamma dose rate
Sensitivity
abstract
Following the Fukushima Daiichi Nuclear Power Plant (FNPP1) accident on March 2011, radioactive
products were released in the atmosphere. Simulations at local scale (within 80 km of FNPP1) were
carried out by the Institute of Radiation Protection and Nuclear Safety (IRSN) with the Gaussian Puff
model pX, during the crisis and since then, to assess the radiological and environmental consequences.
The evolution of atmospheric and ground activity simulated at local scale is presented with a reference
simulation, whose performance is assessed through comparisons with environmental monitoring data
(gamma dose rate and deposition). The results are within a factor of 2e5 of the observations for gamma
dose rates (0.52 and 0.85 for FAC2 and FAC5), and 5e10 for deposition (0.31 for FAC2, 0.73 for FAC5 and
0.90 for FAC10). A sensitivity analysis is also made to highlight the most sensitive parameters. A source
term comparison is made between IRSNs estimation, and those from Katata et al. (2012) and Stohl et al.
(2011). Results are quite sensitive to the source term, but also to wind direction and dispersion pa-
rameters. Dry deposition budget is more sensitive than wet deposition. Gamma dose rates are more
sensitive than deposition, in particular peak values.
Ó2013 Elsevier Ltd. All rights reserved.
1. Introduction
On March 11th 2011, an earthquake of magnitude 9.0 occur-
red off northeastern Japan, causing a tsunami and damaging the
Fukushima Daiichi Nuclear Power Plant (FNPP1). As a result,
radioactive products were released in the atmosphere. During
the emergency phase, the Institute of Radiation Protection and
Nuclear Safety (IRSN) was asked to provide its expertise in
support of the French authorities. Since then, the institute has
been working on improving its assessment of the terrestrial and
marine contamination (Mathieu et al., 2012;Bailly du Bois et al.,
2012). Understanding the formation process of highly con-
taminated areas cannot be achieved through measurements
only. While several kinds of measurements are available, they
only yield partial information: gamma dose rates devices have
a high temporal resolution, but are integrated overall gamma-
emitters, and are too scarce to provide a good spatial coverage.
Soil samplings and airborne readings provide maps of the con-
tamination, but no information on short-lived species, noble
gases, and temporal variations. Thus, improving atmospheric
dispersion simulations remains a key issue, especially for dose
assessments.
*Corresponding author. Tel.: þ33 1 58 35 85 49; fax: þ33146543989.
E-mail address: irene.korsakissok@irsn.fr (I. Korsakissok).
Contents lists available at SciVerse ScienceDirect
Atmospheric Environment
journal homepage: www.elsevier.com/locate/atmosenv
1352-2310/$ esee front matter Ó2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.atmosenv.2013.01.002
Atmospheric Environment 70 (2013) 267e279
To this day, most numerical studies have focused on recon-
structing the source term using environmental data, either at large
scale (Stohl et al., 2011,Winiarek et al., 2012), or at local scale
(Katata et al., 2012), and using it for dose assessments (e.g. Morino
et al., 2011;Terada et al., 2012). Numerical simulations and model-
to-data comparisons at local scale are scarce, partly because of the
difculty to produce satisfactory meteorological elds at that scale.
Such studies (Chino et al., 2011 and following papers) use gamma
dose rates, but not deposition measurements (except Terada et al.,
2012 which is based on prefectural measurements at Japan scale).
The existing studies were conducted with a given source term and
set of deposition parameters, but no extensive sensitivity study has
been carried out.
This paper presents the evolution of atmospheric and ground
activity simulated at local scale (within 80 km of FNPP1). Simu-
lations are made with pX, IRSNs Gaussian puff model. The pX
model is part of the operational platform C3X, which is used by
IRSNs Emergency Response Center in case of an accidental radi-
oactive release. The aim of this study is to better understand the
formation processes of the contaminated areas, but also to give
new insights on the pertinence and limitations of model evalua-
tion tools and indicators in accidental situations. Indeed, usual
Gaussian model evaluations are made on simple, well-known
dispersion experiments. The Fukushima accident provides an un-
precedented case to evaluate atmospheric dispersion models
devoted to radionuclides, with many environmental measure-
ments. We try to highlight advantages and shortcomings of each
kind of measurements and of statistical indicators used to evaluate
a models performance.
More than one year later, many uncertainties remain, especially
on the source term (release kinetics, source height, and isotopic
composition) and meteorology. Besides, uncertainties in simulation
parameters such as dry deposition velocities and scavenging co-
efcients cannot be neglected. Our sensitivity study is aimed at
identifying the most sensitive simulation parameters and input
data.
This paper is organized as follows: rst, input data and simu-
lation set-up are described for a referenceconguration (Section
2). Then, this reference simulation is compared to gamma dose rate
and deposition measurements (Section 3). Finally, the sensitivity
simulations are presented and discussed based on total deposition
budget and model-to-data comparisons (Section 4).
2. Referencesimulation set-up and input data
2.1. Meteorological data
2.1.1. Wind
Operational forecasts from the European Center for Medium-
Range Weather Forecasts (ECMWF) with a spatial resolution of
0.125
0.125
and a time resolution of 3 h were used. The model
does not resolve the complex topography of the Fukushima area,
which is located close to the sea and within 10 km of a mountainous
area. The dataset was therefore analyzed and compared with
available meteorological data at several monitoring stations in the
Fukushima prefecture.
Fig. 1 shows the comparison between the wind at FNPP1 fore-
cast by ECMWF model, and observed by a monitoring car. The
modeled wind speed is often higher than the observed values,
which is consistent with the difference in heights between obser-
vations (monitoring car) and simulation (10 m). Besides, the cell
containing FNPP1 is partly over the ocean, where wind speeds are
globally higher. The wind direction comparison shows a rather
good model-to-data agreement except for some events, especially
during March 15 in the evening. During this period, observations
clearly indicate a northenorthwest plume travel direction, whereas
the modeled direction is mostly west. Accuracy in the wind direc-
tion is an issue of prime importance at that time since it coincides
with heavy rainfalls (cf. Section 2.1.2) and large releases. Therefore,
the simulations used in this study were carried out using homo-
geneous wind elds built with observations at FNPP1 during March
15, between 18 h and midnight, with a 10-min frequency. This so-
lution had its own limitations, since the wind observed at Daiichi
was used on a larger domain than its representativity scale, and did
not take into account vertical wind shear. For the rest of the sim-
ulation, three-dimensional ECMWF data were preferred, to account
for the heterogeneity of the ow.
Fig. 1. Comparison of wind observations (1-h median) at FNPP1 and wind given by ECMWF model, between March 12th and 20th. Left: wind direction (the bands of color indicate
the plume travel direction), right: wind speed. The ECMWF wind is taken in the cell of FNPP1 (without spatial interpolation), at 10 m. (For interpretation of the references to color in
this gure legend, the reader is referred to the web version of this article.)
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279268
2.1.2. Rain
During the accident, several rain episodes occurred. The rst
one took place on 15e16 March and contributed to a signicant
contamination by wet deposition. The second one happened on
20e23 March, both in the Fukushima region and in Tokyo. The
spatial and temporal resolution of the ECMWF model may be
insufcient to accurately represent the spatial and temporal
variability of rain episodes. On the other hand, rain radar obser-
vations provided by the Japan Meteorology Agency
1
have
a higher resolution in time and space but may oversee light and
localized rainfalls. Both datasets were compared to observed
rainfall rates available at several meteorological stations (not
shown). Fig. 2 illustrates that radar data improve the spatial
resolution of rain, even when averaged on the same grid reso-
lution as modeled data. It also shows the tendency of meteoro-
logical forecasts to overestimate the spatial coverage of rainfalls,
due to the coarse resolution. Thus, rain radar data are used in the
reference simulation.
2.2. Source term
The source term used here reconstructs atmospheric releases
occurring between March 12th and 26th, 2011. It includes 73 ra-
dioisotopes, based on the French reactor core inventory, corrected
to account for the nominal reactor power. The methodology used
for this estimation is detailed in other references (Corbin and Denis,
2012;Mathieu et al., 2012;IRSN, 2012).
The main release periods are given in Tab le 1. Two main
release heights were considered. A 20-m height (reactor pressure
vessel) is taken whenever pressure decreases happen in the unit,
and a 120-m height (exhaust stack) is preferred for specic
venting actions. For the hydrogen explosions (events 1 and 3), the
source was assumed to be diluted by the heat and momentum of
the explosion, within 100 m and 300 m on the vertical respec-
tively (Katata et al., 2012). After March 17th (events 7e10),
releases were detected by on-site monitoring devices, but could
not be associated to specic units or events. For these releases,
simulation results were used to infer the source height. For event
9, a height of 50 m was preferred, which is consistent with the
smokes coming from the top of reactor buildings (58 m) at that
time.
2
The radionuclides released in the atmosphere can be classied
into (1) noble gases that neither react with other species, nor de-
posit on the ground, and (2) species that undergo dry and wet
deposition. Fig. 3 shows that noble gases were preponderant
especially in the rst releases (events 1, 2 and 3), since they are
more volatile. The total estimated activity released in the atmo-
sphere was 7.18 10
18
Bq (Becquerel). Noble gases (xenon and
krypton) contributed to most of the released activity (91% of the
activity), iodine to 6%, and cesium to less than 1%. For the major
contributors to gamma dose rate, the estimated released quantity
was 1.97 10
17
Bq for
131
I, 1.68 10
17
Bq for
132
I, 1.08 10
17
Bq for
132
Te, 2.06 10
16
Bq for
137
Cs, 2.78 10
16
Bq for
134
Cs and
5.94 10
18
Bq for
133
Xe. The isotopic ratio
131
I/
137
Cs in the source
term was initially 12 for Unit 1 and about 9 for Units 2 and 3. These
initial values are consistent with the ratio of 10 taken by Chino et al.
(2011), based on concentration measurements in rain water and
vegetation at local scale.
2.3. Dispersion and deposition parameters
IRSNs Gaussian puff model pX was used for the simulations
(Soulhac and Didier, 2008). It handles radioactive decay and decay
products with a comprehensive mechanism. Dispersion in the pX
model is based on a discrete representation of the atmosphere
using PasquilleTurner stability classes (Turner, 1969). In our sim-
ulations, the stability was determined using the temperature gra-
dient between 2 m and 100 m in the meteorological forecasts, for
each cell and time step. The Pasquill standard deviation laws
(Pasquill, 1961) were used in the reference conguration.
Dry deposition and wet scavenging are crucial processes to
model the atmospheric behavior of radionuclides, since the species
deposed on the ground still contribute to the gamma dose-rate
after the plume departure from the area. The deposition prop-
erties of radionuclides strongly depend on three aspects: the gas/
aerosols partitioning, the proportion of organic and inorganic forms
of iodine, and the aerosol size distribution (Sportisse, 2007).
Fig. 2. Maps of rainfall rate on March, 15th at 23 h JST. Left: rain simulated by the meteorological model eRight: radar observations averaged on the same mesh (0.125resolution).
1
http://agora.ex.nii.ac.jp/earthquake/201103-eastjapan/weather/data/radar-
20110311/.
2
http://www.tepco.co.jp/en/press/corp-com/release/11032312-e.html.
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279 269
Without knowledge of these features, simple models were chosen
over detailed microphysical modeling. An apparent deposition ve-
locity of 0.2 cm s
1
is taken for all particles, notably
137
Cs (Brandt
et al., 2002). For iodine, 2/3 of released iodine quantity is
assumed to be gaseous (molecular) and 1/3 in particulate form. The
deposition velocity of molecular iodine I
2
is higher than for par-
ticulate matter: 0.7 cm s
1
(Baklanov and Sørensen, 2001). The
deposition velocities used over sea water are lower than over lands,
namely 0.05 cm s
1
(Pryor et al., 1999). The scavenging coefcient is
given by Ls¼L
0
p
0
B
, with p
0
the rain intensity in mm h
1
. In the
reference simulation, L
0
¼510
5
hmm
1
s
1
(as in Terada et al.,
2012) and B¼1.
3. Reference results and discussion
In this section, simulations with the pX model and the reference
conguration are analyzed and compared with observations. Sta-
tistical indicators used for model-to-data comparisons are dened
in Appendix.
3.1. Gamma dose rates
The observations used in this section come from prefectural
monitoring devices,
3
along with the data provided by TEPCO (Tokyo
Electric Power Company) at Fukushima Nuclear Power Plant 2,
hereafter called Daini.
4
Stations with a low-quality signal (too much
missing data) were discarded. In total, eight monitoring stations
were used (Table 2 and Fig. 7). To compute the simulated gamma
dose rates, 135 radionuclides (including 73 emitted species plus
their decay products) are taken into account. Dose coefcients
(Eckerman and Ryman, 1993) are used to infer dose rates from each
speciesvolume and surface activities. Since the Gaussian model
gives an analytical formula of the concentration, the dose rates
were directly computed at the stationslocations and no simulation
grid was used.
3.1.1. Overall performance on stations
The overall model performance on stations is satisfactory: 52%
of data are within a factor of 2 of the observations (FAC2), and 85%
are within a factor of 5 (FAC5). The fractional bias is 0.44 which
indicates a trend to overestimation, the correlation is 0.72, and the
gure of merit in time (FMT) is 0.43. This compares well to typical
Gaussian models performance on dispersion experiments
(Korsakissok and Mallet, 2009), even though uncertainties on input
data are much higher.
The ambient gamma dose rate measured in air comes from two
contributions: the direct plume contribution (hereafter called
plume-shine) and the gamma-ray emitted by radionuclides
Fig. 3. Time evolution of atmospheric release rate from Fukushima Daiichi Nuclear Power Plant between 12 and 26 March, 2011, for (a) Noble gas and (b) Other species (cesium,
iodine, tellurium.).
Table 1
Release periods of the source term, events that caused the releases (when identied), source height and main plume travel direction.
Number Beginning date Ending date Event Main plume travel direction Source height
1 12/03 10 h 13/03 09 h Unit 1 ehydrogen explosion North, then east 20 m (diluted on 100 m)
2 13/03 08 h 13/03 13 h Unit 3 eventing East (Pacic Ocean) 120 m
3 14/03 05 h 14/03 15 h Unit 3 eventing, then hydrogen explosion East (Pacic Ocean) 150 m (diluted on 300 m)
4 15/03 00 h 15/03 04 h Unit 2 - venting South 120 m
5 15/03 06 h 15/03 21 h Unit 2 ebreach on the wet-well West, north-west, south 20 m
6 16/03 00 h 16/03 15 h Units 2 and/or 3 pressure decreases South 20 m
7 18/03 15 h 18/03 20 h Unit 3? North 120 m
8 20/03 15 h 21/03 04 h Units 2 and/or 3? South 120 m
9 21/03 15 h 23/03 00 h Units 2 and 3 (white and gray smokes) South-west 50 m
10 25/03 08 h 25/03 10 h Unit 2? West 120 m
3
http://www.pref.fukushima.jp/j/7houbu0311-0331.pdf and http://www.pref.
fukushima.jp/j/20-50km0315-0331.pdf.
4
http://www.tepco.co.jp/en/nu/fukushima-np/f2/data/2011/index-e.html.
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279270
deposed on the ground (ground-shine). The plume-shine usually
is responsible for peak values (high but during a short time),
whereas the ground-shine corresponds to the gamma dose-rate
measured after the plume departure. Hence, the most numerous
gamma-dose rate observations correspond to ground-shine, which
decreases slowly due to radioactive decay. The decrease rate of
this residual gamma dose rate depends on the radioisotopes
deposed on the ground, and on their respective half-life times.
Therefore, classicalindicators such as FAC2 and FAC5 mainly
depend on the simulations ability to forecast deposition, isotopic
composition and subsequent decay. In case of an accidental release
of radionuclides, an operational simulation should be able to
forecast peak values, since they represent an important part of
human exposure (through direct radiation and inhalation), and
plume arrival times (i.e. the rst date when the gamma dose rate
value is higher than background value). In the following, we will
focus on (1) bias on peak values, (2) plume arrival times and (3)
FAC2, FAC5 and FMT. Comparisons are made on hourly-averaged
values.
Table 2 gives an overview of the models performance for each
station. Peak values are within less than a factor of two, except for
Iwaki, Daini and Minamisoma, where it is overestimated by a fac-
tor of ve. These three stations are located close to the coast,
where the meteorological model often has difculties forecasting
the wind eld. Besides, for the two events responsible for the
peaks (event 1 for Minamisoma and event 4 for Daini and Iwaki),
the meteorological conditions were very stable. Thus, the plume
was very thin, and uncertainties in the wind eld, station location,
and/or release height would have a large impact on the result. At
Minamisoma and Iwaki, gamma dose-rate measurements are only
available every hour during the main releases periods (prior to
March 16). Simulations show a high temporal variability, especially
during the plume passage, indicating that the temporal frequency
of observations may not be sufcient. The temporal resolution of
the wind eld (3 h) is also too coarse to account for the wind
variability.
On the other stations, the peak values are very well repro-
duced, although a delay of 6 h in the plume arrival time is
observed on the northwestern stations Iitate and Fukushima. At
Funehiki, there were no observations at the simulated peak
times, and the air dose-rate due to deposition is overestimated
by more than a factor of ve. For all other stations except Iwaki,
more than 80% of simulated values are within a factor 5 of the
observations, and the FAC2 is also very good, especially at the
northwestern stations. Fig. 4 shows the model-to-data compari-
sons of gamma dose rate values, for the eight monitoring sta-
tions, hour-by-hour. A few singular values are underestimated,
especially at Iitate and Fukushima, due to the delay in the plume
arrival time.
3.1.2. Temporal analysis on stations
Fig. 5 shows the temporal evolution on four of these stations:
two are representative of a high contamination level due to wet
deposition (Iitate and Fukushima), and two stations where little
wet deposition occurs (Minamisoma and Kawauchi). The latter
show a time series with several peaks corresponding to plumes
coming through the station, and little deposition, while the former
show a high contamination by deposition, with a decrease rate
over time essentially due to radioactive decay of the deposed
isotopes.
At Minamisoma and Kawauchi (Fig. 5(a) and (b)), most peaks
are simulated, with a small delay and some overestimation at
Kawauchi. Missing peaks are probably due to inaccuracies in the
source term, but are small compared to the initial con-
tamination. These two stations illustrate the difculty to cor-
rectly represent both peak values and deposition: at
Minamisoma, the peak value is overestimated but the deposition
is correctly forecast, while at Kawauchi, the peak value is much
better reproduced but the deposition is overestimated after
March 22nd. At Iitate and Fukushima, the dose rates are well
reproduced, except for the initial delay in the plume arrival (6 h).
This delay is also found by other simulations (Katata et al., 2012).
Fig. 6 illustrates the contamination processes at two typical
Table 2
Comparison of observed and simulated ambient air dose rate at the eight stations: peak time and value (the peak is the maximum value on the whole period of observations),
Fac2 and Fac5 (proportion of simulation values within a factor of 2 resp. 5 of the observations). Statistics are made on the whole period (12/03e26/03).
Station name Event(s) associated
with peaks
Observed arrival time
(1st peak)
Simulated arrival
time (1st peak)
Observed peak
value (
m
Gy h
1
)
Simulated peak
value (
m
Gy h
1
)
Fac2 (%) Fac5 (%) FMT
Kawauchi Event 5; event 9 15/03 11:00 15/03 12:00 11.5 15.7 69 99 0.64
Tamura City
Funehiki
Event 5 (2 peaks) e15/03 16:00
16/03 06:00
e6.2
9.4
5 82 0.19
Koriyama Event 5; event 9 15/03 14:00 15/03 17:00 6 3.0 52 99 0.47
Iitate Event 5 15/03 15:00 15/03 21:00 39.5 57.6 67 97 0.54
Fukushima Event 5 15/03 16:00 15/03 22:00 24 19.1 92 98 0.79
Minamisoma Event 1; event 7 12/03 20:00 12/03 19:00 20.0 87 95 100 0.68
Daini (FNPP2) Event 4; event 6 15/03 00:00 15/03 01:00 94 515 12 88 0.38
Iwaki Event 4; event 6 15/03 01:00 15/03 03:00 23.7 147 9 16 0.12
Fig. 4. Scatter plot of gamma dose rate values (
m
Gy h
1
) at the eight monitoring sta-
tions for the reference simulation. Red line: perfect agreement. Dashed line (bold):
factor 2. Dashed line: factor 5.
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279 271
stations: (a) Kawauchi, where the plume contribution is pre-
ponderant (91% of the peak gamma dose rate) and ground shine
is only due to dry deposition, and (b) Fukushima, where the wet
deposition contribution to the gamma air dose rate is predomi-
nant (95% of the peak dose rate). Fig. 6(b) shows that the delay at
Fukushima is not due to inaccuracies in the rain timing, but to
a delay in the plume arrival time. Indeed, the plume contribution
does not increase signicantly before 22 h on March 15th. This
might be due to uncertainties in the wind eld and/or source
height.
Fig. 5. Comparison of observed and simulated gamma air dose rate at four of the eight monitoring stations. Simulations are made with pX, for the reference conguration.
Fig. 6. Plume, dry deposition and wet deposition contributions to the simulated ambient air gamma dose rate at (a) Kawauchi and (b) Iitate.
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279272
3.2. Deposition
The simulations used in this section are made on a polar mesh,
with circles at 2.5, 5, 10, 15, 20, 25, 30, 40, 50, 65 and 80 km from
FNPP1, and a ten-degree step for the angle.
3.2.1. Spatial analysis
Fig. 7 shows the simulated spatial distribution of deposition
after the end of releases (March 30, 2011). Dry deposition mostly
occurred along the coast, north and south from FNPP1 ((Fig. 7(a)),
whereas wet deposition strongly contaminated a northwestern
area, including Iitate and Fukushima (Fig. 7(b)). Analyses by event
showed that the dry deposition north of FNPP1 was formed
during two distinct events: event 1 (12/03) and event 7 (18/03).
The southern deposition along the coast corresponds to event 4
(15/03 at 00 h) and event 6 (16/03). Wet deposition mainly cor-
responds to event 5 (67% of total land deposition occurred within
few hours on March 15th), but also occurred on 21/03 and 22/03
(16% of total deposition), south and south-west of the plant
(event 9). The dry deposition over the sea is much lower, due to
the use of lower deposition velocities over water. The con-
tribution of dry deposition to total deposition is shown Fig. 7(c).
It conrms that the northwestern contamination is mainly due to
wet deposition (more than 90% of the total deposition), whereas
dry deposition contributes to almost 100% deposition along the
coast.
Dry and wet deposition patterns are similar for other species
than
137
Cs, although dry deposition is stronger for molecular iodine.
This leads to a higher ratio
131
I/
137
Cs where dry deposition is pre-
dominant, especially south of FNPP1 (Fig. 7(d)). This pattern is
consistent with observations (Kinoshita et al., 2011). This ratio
tends to decrease over time, since the half-life of
131
I (eight days) is
much lower than for
137
Cs (thirty years).
3.2.2. Deposition budget
As Morino et al. (2011) did on Japan scale, we computed the total
budget of
137
Cs and
131
I deposed on the simulation domain through
dry and wet deposition (decay-corrected). Their source term comes
from Chino et al. (2011), which includes 13 PBq of
137
Cs (20.6 PBq in
our estimation) and 150 PBq of
131
I (197 PBq in our case). Morino
et al. (2011) supposed that the gaseous fraction of
131
I was 80% of
the total release (67% in our case). Their estimation for the
Fukushima prefecture can be compared with our land deposition
budget within 80 km from FNPP1 (Table 3).
The absolute values are of the same order of magnitude, except
for
137
Cs dry deposition, where our value is ten times higher. This
Fig. 7. Spatial distribution of (a)
137
Cs dry deposition, (b)
137
Cs wet deposition, (c) Ratio of dry deposition on total deposition for
137
Cs and (d)
131
I/
137
Cs ground activity ratio. Values
are given on March 30, 2011.
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279 273
may be due to the lower released amount, and to lower dry dep-
osition velocities (0.1 cm s
1
instead of 0.2 cm s
1
in our simula-
tions). For gaseous
131
I, Morino et al. (2011) took 0.5 cm s
1
(0.7 cm s
1
in our case). Estimates of total iodine dry deposition are
comparable. Wet deposition values are more difcult to compare,
without knowledge of the parameterization used. In our case,
scavenging is the same for all species. In Morino et al. (2011), the
wet deposition scheme seems more efcient for
137
Cs, which is in
particulate form, than for
131
I, which is mostly gaseous. It probably
explains the higher proportion of
137
Cs deposed over land (12% of
their release).
Total deposition values are very close to each other. Compen-
sation between dry and wet deposition partly explains this agree-
ment. Based on airborne monitoring data
5
, the total quantity of
137
Cs within 80 km of the plant over land was estimated to be about
1.49 PBq, with an uncertainty of 0.49 PBq (Gonze, M-A, personal
communication). Both simulations are within this range.
3.2.3. Comparisons to deposition measurements
Fig. 8 shows the comparison between the simulation and
the observations of
137
Cs deposition provided by Ministry of
Education, Culture, Sports, Science and Technology (MEXT).
6
About 1800 measurements are within our simulation domain.
The overall shape of the northwestern contamination (over
10
5
Bq m
2
) is correct, but the highest values are located too
north compared to the observations. This is probably due to the
use of wind observations at FNPP1 at that time, not representa-
tive of the wind direction at a larger scale. Thus, the model
overestimates deposition north along the coast by more than
a factor ve (yellow to red points, Fig. 8(c)). The overestimation
south, after 40 km, is consistent with gamma dose rates at Iwaki.
Fig. 8(d) shows that the agreement is better between 20 and
60 km from the source. Closer to the source, the model tends to
overestimate deposition. Between 60 and 80 km, values are
underestimated, especially in the northern area (gray and purple
points, Fig. 8(c)). In all, 31% of simulated values are within a factor
2 of the observations, 73% within a factor 5, and 90% within
a factor 10. The correlation coefcient is 0.34. The gure of merit
in space (FMS) depends on the chosen threshold. A high
threshold would focus on the models ability to forecast extreme
values. For 10
4
Bq m
2
(94% of measurements), the FMS is very
good (0.85). For 10
5
Bq m
2
(30% of measurements, mainly
northwest), it goes down to 0.43, probably because of the mis-
placed wet deposition zone.
Very close measurement points may differ by a factor 5e10,
which may be due to running water or changes in terrain type
(school yard, agricultural eld.). The simulation gives averaged
values, and does not account for local-scale variability. Thus, fur-
ther analyses of the datasets and aggregating close measurements
could improve these comparisons.
4. Sensitivity study
4.1. Sensitivity parameters
Table 4 shows an overview of the parameters chosen for the
sensitivity simulations. Simulations were carried out independ-
ently for each parameter, to assess their impact on results.
For dispersion, several Gaussian standard deviation formulations
were used. For dry and wet deposition, values were taken within the
range of the literature, i.e. more or less within a decade (Sportisse,
2007). Mixing the release between 0 and 150 m allowed high-
lighting the sensitivity to the release height. Several source terms
were also compared. The release from Katata et al. (2012) was
designed at local scale, including several isotopes to compute the
gamma dose rates. Thus, it was included in deposition and gamma
dose rate sensitivity runs. The release from Stohl et al. (2011), cali-
brated on long-range simulations (with a rst-guess), did not
include short-lived species that are major contributors to gamma
dose rate. Thus, it was only used in the comparisons for
137
Cs dep-
osition. Another IRSNs estimation, using inverse modeling with
gamma dose rate observations (Saunier et al., 2012), was included.
Concerning meteorological data, the inuence of rain (radar obser-
vations versus rain forecasts) and wind elds (with and without
wind observations at Daiichi on March 15th) was evaluated.
4.2. Sensitivity results
4.2.1. Sensitivity of total deposition budget of
137
Cs
Fig. 9 illustratesthe sensitivity of the deposition budget overland.
The sensitivity of total deposition is conditioned by that of wet
deposition, whichaccounts for 2/3 of total deposition (cf. Table3). For
total deposition, most simulations are within a factor 2 of the refer-
ence simulation, except with the source term from Stohl et al. (2011)
which has a higher estimation of
137
Cs release (35 PBq of
137
Cs). Ac-
cording to airborne observations, the total land deposition within
80 km of FNPP1 should be between 1 and 2 PBq (see Section 3.2.2).
Most simulations meet this condition except the Release_Stohl and
lmin congurations. Dry deposition (Fig. 9(a)) is also sensitive to
vertical diffusion. Indeed, Briggs urban and constant diffusion pa-
rameterizations enhance vertical diffusion, which induces lower dry
deposition, since the plume is less concentrated near the ground.
Finally, a compensation mechanism between dry andwet deposition
appears: lower deposition velocities (vdmin) imply that the plume is
less depleted near the source. Hence, the plume wash-out by the rain
is more efcient, and wet deposition increases (Fig. 9(b)).
Table 3
Simulated
137
Cs and
131
I loss due to dry and wet deposition: contribution over land and sea, and total on the simulation domain (80 km around FNPP1). (*): deposed quantity
normalized by released quantity. Simulations are made with pX, for the reference conguration. Values are given in PBq (10
15
Bq). The values are decay-corrected for iodine.
Morino et al., 2011
Fukushima pref. 13,783 km
2
Land 80 km of FNPP1
about 10
4
km
2
Sea 80 km of FNPP1
about 10
4
km
2
Total (land þsea,
80 km radius)
131
I (PBq) Decay-
corrected
Dry deposition 8.07 7.9 0.3 8.2
Wet deposition 3.6 6.1 22.0 28.1
Total deposition 12.3 (8.2%)* 14.0 (7.1%)* 22.3 36.3 (18.4%)*
137
Cs (PBq) Dry deposition 0.047 0.39 0.02 0.39
Wet deposition 1.48 0.94 2.6 3.9
Total deposition 1.53 (11.7%)* 1.33 (6.4%)* 2.62 4.29 (20.8%)*
5
http://www.mext.go.jp/english/incident/1304796.htm.
6
http://www.mext.go.jp/b_menu/shingi/chousa/gijyutu/017/shiryo/__icsFiles/
aeldle/2011/09/02/1310688_1.pdf.
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279274
4.2.2. Sensitivity of spatial deposition of
137
Cs
The sensitivity of spatial deposition may be investigated
through statistical indicators used in comparison to MEXT mea-
surements (Fig. 10). They are quite robust to changes in parame-
terizations such as standard deviations, vertical mixing, and dry
deposition. The source term is the most sensitive, along with rain
and wind elds. The Release_Stohl decreases signicantly the FAC2,
and changes in meteorological data also reduce the FAC2
(Fig. 10(a)). A low scavenging coefcient (lmin) increases FAC2
(0.37), but lowers FAC5 (0.70). FAC2 and FAC5 would not penalize
a simulation that would homogenize the deposition. For instance,
asimulatedhomogeneous deposition of 10
5
Bq m
2
, compared to
Fig. 8. Comparison of observed and simulated deposition values for
137
Cs. (a) Observations from MEXT. (b) Map of bias factor (observation/simulation). (c) Deposition simulated
with pX. (d) Scatter plot: Red line: perfect agreement. Dashed line (bold): factor 2. Dashed line: factor 5.
Table 4
Sensitivity simulations: overview of all simulation parameters modied in the sensitivity runs. Each parameter is made to vary independently from one another. For each
sensitivity parameter, the name used in the gures is given in brackets.
Parameter Reference Value 1 (name) Value 2 (name) Value 3 (name)
Standard deviations Pasquill Briggs rural (briggs_rural) Briggs urban (briggs_urban) Constant diffusion K (diff)
Dry deposition particles
(iodine)
210
3
(7 10
3
)ms
1
510
4
(1 10
3
)ms
1
(vdmin)510
3
(2 10
2
)ms
1
(vdmax)
Wet deposition 5 10
5
hmm
1
s
1
110
5
(lmin)110
4
(lmax)
Release height Time varying (cf Table 1) Diluted between 0 and 150 m
(Vertical_mixing)
Source term Mathieu et al., 2012 Katata et al., 2012 (Release_Chino)Stohl et al., 2011 (Release_Stohl)Saunier et al., 2012
(Release_Inverse)
Rain Rain radar observations ECMWF forecasts (ECMWF_rain)
Wind ECMWF þobservations
at FNPP1
ECMWF forecasts (ECMWF_wind)
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279 275
observations, would have 0.33 as FAC2 and 0.73 as FAC5, which is
within the other simulationsresults. Thus, other indicators such as
FMS (Fig. 10(b)) and deposition maps (Supplementary material)are
used as a complement.
The FMS is compared for a 10
5
Bq m
2
threshold which cor-
responds to the northwestern deposition. It is therefore sensitive
to wet deposition parameters (lmax,lmin) and to the wind used
during March 15th (ECMWF_wind). The Release_Stohl tends to
overestimate deposition in the northwestern area, but is not
penalized by this indicator, and the Release_Chino improves the
FMS.
4.2.3. Sensitivity of gamma dose rates
Gamma dose rates are much more sensitive than deposition,
which is integrated in time and has a better spatial coverage.
With only eight stations, local discrepancies between model and
reality on one station may have a huge impact on the results.
Thus, a simulation may have acceptable results on deposition,
but not on gamma dose rates. This is the case of briggs_urban
vertical_mixing and Release_inverse simulations (Fig. 11). Lower
dry deposition velocities improve the FMT by compensating for
the overestimation due to errors in the source term and/or
meteorological data. Dispersion parameters and source height
have a much larger inuence on gamma dose rates, especially
peak values, than on deposition. This is particularly true on
coastal stations, where the plume is very thin and the peak
values are very sensitive to changes in the wind direction and
diffusion. For instance, using a constant diffusion coefcient
(diff) improves the results on coastal stations by increasing the
plume vertical dilution in stable situations (Fig. 12(a) and (c)),
thus reducing the overestimation. Using ECMWF forecasts
without observations at FNPP1 (ECMWF_wind) drastically lowers
the model performance on stations because of errors in the
westerly wind direction on March 15th: Iitate and Fukushima are
missed by the plume, and the peak at Koriyama is greatly
overestimated (Fig. 12(b) and (d)). Fig. 12 illustrates the difculty
Fig. 9. Sensitivity of total
137
Cs deposition budget over land, within 80 km of FNPP1. Values for the reference simulation (similar to Table 3) are given in red (dashed lines). (For
interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279276
to have a good model-to-data agreement (i.e. less than a factor 5)
on all stations: the Release_inverse simulation greatly improves
the results on coastal station but overestimates the northwest-
ern peaks, while the Release_chino is better on northwestern
stations but underestimates the peaks on southern and western
stations.
5. Conclusions and perspectives
We presented atmospheric dispersion simulations of the
Fukushima Nuclear Power Plant accident, using IRSNs Gaussian
puff model pX. The evolution of atmospheric and ground activity
simulated at local scale was presented with a referencesimu-
lation, whose performance was assessed through comparisons with
environmental monitoring data (gamma dose rate and deposition).
The results were within a factor of 2e5 of the observations for
gamma dose rates, and 5e10 for deposition. The total quantity of
137
Cs deposed over land and the isotopic ratio
131
I/
137
Cs were
consistent with observations and with other estimations. The
gamma dose rates in the northwestern area, highly contaminated
by wet deposition, were correctly reproduced but the wet
deposition zone was slightly misplaced. The coastal stations were
more difcult to simulate, due to the stable situation inducing
a thin plume, hence a large dependency to uncertainties in the
wind direction and station location. Besides, the temporal
frequency of measurements and/or wind elds may not be
sufcient to correctly reproduce the peak values when the plume
passed through the stations in less than 1 h. Thus, it is difcult to
determine whether the overestimation along the coast is due to
the source term, meteorological data and/or diffusion parameters.
Fig. 10. Sensitivity of comparisons to MEXT measurements of
137
Cs, within 80 km of FNPP1. Values for the reference simulation are given in red (dashed line). (For interpretation of
the references to color in this gure legend, the reader is referred to the web version of this article.)
Fig. 11. Sensitivity of comparisons to gamma dose rate measurements on the eight monitoring stations. Values for the reference simulation are given in red (dashed line). (For
interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279 277
While uncertainties in the input data (wind direction, rain,
source term kinetics and quantity) are huge, simulation parameters
are also uncertain. The inuence of each parameter was assessed
separately. As expected, source terms had the highest impact on the
results: the release from Stohl et al. (2011) tended to over-
estimation; that from Katata et al. (2012) gave better results on the
northwest but underestimation in the south and west. Total dep-
osition budget was within a factor 2 of the reference most of the
time. The sensitivity study seemed to validate our choice for scav-
enging coefcients. Gamma dose rates, especially peak values, were
quite sensitive: source terms, standard deviations, and wind di-
rection had huge impacts on the results. However, with only eight
monitoring stations, each station has a particular behavior and no
general conclusions can be made.
This study is a rst step toward updating model evaluation
tools and proposing new indicators for modeling accidental re-
leases of radionuclides. Despite the situation complexity and the
remaining uncertainties, the results presented here are satisfactory
compared to standard dispersion model evaluations. They can
therefore be used as a benchmark for atmospheric dispersion
modelsevaluation and improvement with respect to emergency
purposes. The main perspective is to use ensemble simulations
instead of a single, deterministic model, in order to account for
uncertainties in the input data and simulation parameters (Mallet
and Sportisse, 2008). Another idea is to account for uncertainties
in the plume position by comparing measurements to a cloud of
points, or to an average over a given volume, instead of a single
value.
Acknowledgments
We want to express our sympathy to Japanese people who
endured the terrible consequences of the earthquake, tsunami and
nuclear accident.
We thank MEXT and TEPCO for the online publication of
measurements.
We thank Météo France for providing meteorological data, D.
Corbin and J. Denis for the source term estimation, R. Gurriaran for
Fig. 12. Sensitivity of gamma dose rate peak values on four monitoring stations. Values for the reference simulation are given in red (dashed line). Observation values are given in
black (bold dashed line). (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279278
his expertise on measurements, and people from the Emergency
Response Center who worked during the crisis.
Appendix A. Statistical indicators
We consider a set of measurements O
i
(ibetween 1 and N) and
the corresponding model outputs P
i
. Statistical indicators are
dened as follows (Chang and Hanna, 2004):
(1) Fractional bias MFBE:
MFBE ¼X
N
i¼0
P
i
O
i
P
i
þO
i
(2) Bias Factor at a given point:
BF ¼P
i
O
i
(3) Correlation coefcient:
r¼P
N
i¼1
O
i
OP
i
P
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P
N
i¼1
O
i
O
2
P
N
i¼1
P
i
P
2
r
(4) FAC2 (resp. FAC5): proportion of values such that 0.5 BF 2
(resp. 0.2 BF 5).
(5) The Figure of Merit in Time (FMT) is the percentage of overlap
between measured and predicted integrated time series at
a given location:
FMT ¼P
N
i¼1
minðO
i
;P
i
Þ
P
N
i¼1
maxðO
i
;P
i
Þ
(6) The Figure of Merit in Space (FMS) is the percentage of overlap
between measured and predicted areas above a threshold Tat
a given time. A
P
is the number of points verifying P
i
Tand A
O
is the number of points such that O
i
T, then
FMS ¼A
P
XA
O
A
P
WA
O
Aperfectmodel-to-data agreement corresponds to MFBE ¼0
and other indicators equal to 1.
Appendix B. Supplementary material
Supplementary material related to this article can be found at
http://dx.doi.org/10.1016/j.atmosenv.2013.01.002.
References
Bailly du Bois, P., Laguionie, P., Boust, D., Korsakissok, I., Didier, D., 2012. Estimation
of marine source-term following Fukushima Daiichi accident. Journal of Envi-
ronmental Radioactivity 114, 2e9.
Baklanov, A., Sørensen, J.H., 2001. Parameterization of radionuclide deposition in
atmospheric long-range transport modelling. Physics and Chemistry of the
Earth (B) 26, 787e799.
Brandt, J., Christensen, J.H., Frohn, L., 2002. Modelling transport and deposition of
caesium and iodine from the Chernobyl accident using the DREAM model.
Atmospheric Chemistry and Physics 2, 397e417.
Chang, J.C., Hanna, S.R., 2004. Air quality model performance evaluation. Meteor-
ology and Atmospheric Physics 87, 167e196.
Chino, M., Nakayama, H., Nagai, H., Terada, H., Katata, G., Yamazawa, H., 2011.
Preliminary estimation of release amounts of
131
I and
137
Cs accidentally dis-
charged from the Fukushima Daiichi Nuclear Power Plant into the atmosphere.
Journal of Nuclear Science and Technology 48, 1129e1134.
Corbin, D., Denis, J., 2012. Evaluation des rejets atmosphériques liés à laccident de
Fukushima. IRSN. Report PSN-RES/SAG/2012e00347.
Eckerman, K.F., Ryman, J.C., 1993. Exposure-to-dose Coefcients for General
Application, Based on the 1987 Federal Radiation Protection Guidance. Oak
Ridge National Laboratory. Report 12.
IRSN, 2012. Fukushima, One Year Later: Initial Analyses of the Accident and
Its Consequences. IRSN. Report IRSN/DG/2012-003. http://www.irsn.fr/EN/
publications/technical-publications/Documents/IRSN_Fukushima-1-year-later_
2012-003.pdf.
Katata, G., Ota, M., Terada, H., Chino, M., Nagai, H., 2012. Atmospheric discharge
and dispersion of radionuclides during the Fukushima Daiichi Nuclear Power
Plant accident. Part I: source term estimation and local-scale atmospheric
dispersion in early phase of the accident. Journal of Environmental Radio-
activity 109, 103e113.
Kinoshita, N., Sueki, K., Sasa, K., Kitagawa, J.-i., Ikarashi, S., Nishimura, T., Wong, Y.-S.,
Satou, Y., Handa, K., Takahashi, T., Sato, M., Yamagata, T., 2011. Assessment of
individual radionuclide distributions from the Fukushima nuclear accident
covering central-east Japan. Proceedings of the National Academy of Sciences
108 (49), 19526e19529.
Korsakissok, I., Mallet, V., 2009. Comparative study of Gaussian dispersion formulas
within the Polyphemus platform: evaluation with Prairie Grass and Kincaid
experiments. Journal of Applied Meteorology 48, 2459e2473.
Mallet, V., Sportisse, B., 2008. Air quality modeling: from deterministic to stochastic
approaches. Computers & Mathematics with Applications 55, 2329e2337.
Mathieu, A., Korsakissok, I., Quélo, D., Groëll, J., Tombette, M., Didier, D., Quentric, E.,
Saunier, O., Benoit, J.-P., Isnard, O., 2012. Atmospheric dispersion and deposition
of radionuclides from the Fukushima Daiichi nuclear power plant accident.
Elements 8, 195e200.
Morino, Y., Ohara, T., Nishizawa, M., 2011. Atmospheric behavior, deposition, and
budget of radioactive materials from the Fukushima Daiichi nuclear power
plant in March 2011. Geophysical Research Letters 38, L00G11.
Pasquill, F.A., 1961. The estimation of the dispersion of windborne material. The
Meteorological Magazine 90 (1063), 33e49.
Pryor, S.C., Barthelmie, R.J., Geernaert, L.L.S., Ellermann, T., Perry, K.D., 1999.
Speciated particle dry deposition to the sea surface: results from ASEPS 97.
Atmospheric Environment 33, 2045e2058.
Saunier, O., Mathieu, A., Didier, D., Tombette, M., Quélo, D., Winiarek, V.,
Bocquet, M., 2012. Using gamma dose rate monitoring with inverse modelling
techniques to estimate the atmospheric release of a nuclear power plant acci-
dent: application to the Fukushima case. In: International Meeting on Severe
Accident Assessment and Management: Lessons Learned from Fukushima Dai-
ichi., San Diego, USA, 11e15 November 2012.
Soulhac, L., Didier, D., 2008. Projet pX, note de principe pX 1.0. IRSN. Report IRSN/
DEI/SESUC/08-39.
Sportisse, B., 2007. A review of parameterization for modelling dry deposition and
scavenging of radionuclides. Atmospheric Environment 41, 2683e2696.
Stohl, A., Seibert, P., Wotawa, G., Arnold, D., Burkhart, J.F., Eckhardt, S., Tapia, C.,
Vargas, A., Yasunari, T.J., 2011. Xenon-133 and caesium-137 releases into the
atmosphere from the Fukushima Dai-ichi nuclear power plant: determination
of the source term, atmospheric dispersion, and deposition. Atmospheric
Chemistry and Physics Discussions 11, 28319e28394.
Terada, H., Katata, G., Chino, M., Nagai, H., 2012. Atmospheric discharge and dis-
persion of radionuclides during the Fukushima Dai-ichi Nuclear Power Plant
accident. Part II: verication of the source term and analysis of regional-scale
atmospheric dispersion. Journal of Environmental Radioactivity 112, 141e154.
Turner, D.B., 1969. Workbook of Atmospheric Dispersion Estimates.
Winiarek, V., Bocquet, M., Saunier, O., Mathieu, A., 2012. Estimation of errors in the
inverse modeling of accidental release of atmospheric pollutant: application to
the reconstruction of the cesium-137 and iodine-131 source terms from the
Fukushima Daiichi power plant. Journal of Geophysical Research 117, D05122.
I. Korsakissok et al. / Atmospheric Environment 70 (2013) 267e279 279
... Because of the sparsity of monitoring sites, the atmospheric dispersion model becomes an inevitable tool to determine the spatiotemporal distribution of atmospheric radionuclides and to assess the accident consequence. And the local-scale dispersion within 80 km is of particular interest for the early stages after a nuclear accident (Korsakissok et al., 2013) and source term estimation (Katata et al., 2015;Terada et al., 2020a). ...
... Until now, numerous atmospheric dispersion models have been developed for the local-scale dispersion, including Gaussian, Lagrangian, and Eulerian models (Connan et al., 2013;Dong et al., 2021a,b,c;Korsakissok et al., 2013;Liu et al., 2017;Sekiyama and Kajino, 2021;Wang et al., 2020). However, most studies of Fukushima accident atmospheric dispersion focus on a larger scale (>80 km) (Kadowaki et al., 2021;Katata et al., 2012Katata et al., , 2015Park et al., 2013;Sato et al., 2020;Terada et al., 2020a). ...
... This study revealed that the increase of resolution may not reproduce a detailed local-scale atmospheric dispersion process, because of the artificial numerical diffusion effect of the Eulerian method (Sekiyama and Kajino, 2021). Another study focused on an area within 80 km of the FDNPP1 to evaluate the performance of the Gaussian Puff model pX at eight gamma dose rate monitoring stations and identified the critical factors that affected significantly the simulation of the pX model (Korsakissok et al., 2013). ...
... The dispersion of radionuclides released into the atmosphere depends on the physicochemical properties of the released substances, the emission parameters (e.g., source elevation, timing, and duration of the release), and meteorological conditions at the accident site (e.g., wind speed and direction) (Girard et al., 2014). In order to forecast the dispersion of radionuclides during the early phase of nuclear accidents and to support decisions and warnings, atmospheric dispersion models (ADMs) are commonly used to predict the transport of radioactive pollutants through the atmosphere as well as the quantities of radioactive material deposited on the ground (Korsakissok et al., 2013). This information is essential for decision makers in order to anticipate the countermeasures necessary to protect the population against contamination. ...
... The main purpose of the present article is to investigate the impact of the meteorological uncertainties on local-scale dispersion. The operational high-resolution meteorological ensembles AROME-EPS (Applications of Research to Operations at Mesoscale-Ensemble Prediction System) (Bouttier et al., 2012) and AROME-deterministic NWP (Seity et al., 2011) of Météo-France are used as input of the IRSN shortrange Gaussian puff model pX (Soulhac and Didier, 2008;Mathieu et al., 2012;Korsakissok et al., 2013) around the RP facility at local scales (less than 20 km). In this area, there is a dense weather observation network (from both IRSN and Météo-France) that has been used to validate AROME-EPS ensembles before combining them with the dispersion model. ...
... The IRSN's Gaussian puff model, pX, used in this work is part of the operational platform C3X (Tombette et al., 2014), which is used by IRSN Emergency Response Center in case of an accidental radioactive release. The pX model is used to simulate the atmospheric dispersion of radionuclides on short and medium distances (500 m-50 km) (Korsakissok et al., 2013;Mathieu et al., 2012). The principle of such a dispersion model is based on the following assumptions: ...
Article
Full-text available
Numerical atmospheric dispersion models (ADMs) are used for predicting the health and environmental consequences of nuclear accidents in order to anticipate countermeasures necessary to protect the populations. However, these simulations suffer from significant uncertainties, arising in particular from input data: weather conditions and source term. Meteorological ensembles are already used operationally to characterize uncertainties in weather predictions. Combined with dispersion models, these ensembles produce different scenarios of radionuclide dispersion, called “members”, representative of the variety of possible forecasts. In this study, the fine-scale operational weather ensemble AROME-EPS (Applications of Research to Operations at Mesoscale-Ensemble Prediction System) from Météo-France is coupled with the Gaussian puff model pX developed by the IRSN (French Institute for Radiation Protection and Nuclear Safety). The source term data are provided at 10 min resolution by the Orano La Hague reprocessing plant (RP) that regularly discharges 85Kr during the spent nuclear fuel reprocessing process. In addition, a continuous measurement campaign of 85Kr air concentration was recently conducted by the Laboratory of Radioecology in Cherbourg (LRC) of the IRSN, within 20 km of the RP in the North-Cotentin peninsula, and is used for model evaluation. This paper presents a probabilistic approach to study the meteorological uncertainties in dispersion simulations at local and medium distances (2–20 km). First, the quality of AROME-EPS forecasts is confirmed by comparison with observations from both Météo-France and the IRSN. Then, the probabilistic performance of the atmospheric dispersion simulations was evaluated by comparison to the 85Kr measurements carried out during a period of 2 months, using two probabilistic scores: relative operating characteristic (ROC) curves and Peirce skill score (PSS). The sensitivity of dispersion results to the method used for the calculation of atmospheric stability and associated Gaussian dispersion standard deviations is also discussed. A desirable feature for a model used in emergency response is the ability to correctly predict exceedance of a given value (for instance, a dose guide level). When using an ensemble of simulations, the “decision threshold” is the number of members predicting an event above which this event should be considered probable. In the case of the 16-member dispersion ensemble used here, the optimal decision threshold was found to be 3 members, above which the ensemble better predicts the observed peaks than the deterministic simulation. These results highlight the added value of ensemble forecasts compared to a single deterministic one and their potential interest in the decision process during crisis situations.
... Because of their spatiotemporal sparsity, ground-based observations alone might not capture all the transport processes. In this case, atmospheric transport modeling represents a key tool for analyzing radionuclide transport behavior and for predicting the environmental consequences (Korsakissok et al., 2013). ...
... For example, there are only a few Tokyo Electric Power Company (TEPCO) monitoring sites and one automated meteorological data acquisition system site within 20 km of the FDNPP, and 99.5% of the critical suspended particulate matter observations were acquired more than 20 km away from the FDNPP (Tsuruta et al., 2014). Direct local simulation of the area within 20 km of the FDNPP also confronts a series of technological challenges, e.g., reproduction of the frequently varying meteorological fields (Korsakissok et al., 2013) and modeling of the complex cross-scale transport behaviors at the site and up to 20 km range. ...
Article
Modeling of local atmospheric radionuclide transport is essential for nuclear emergency response. However, very few studies of the Fukushima Dai-ichi nuclear power plant (FDNPP) accident have focused on this topic because of the complex meteorology and the cross-scale transport behaviors from the site to 20 km of the FDNPP. In this study, both the local meteorology and transport behaviors were investigated at high resolution (200 m) using ensembles of different meteorology and models. Four wind fields calculated from onsite observations and three regional-scale meteorological fields (i.e., the 1-km ECMWF, 3-km and 1-km NHM-LETKF), and two transport models: the RIMPUFF Lagrangian puff model and the SPRAY particle model were considered and combined with each other. These eight simulations and their ensemble mean were analyzed based on onsite observations of wind and gamma dose rates, and local-scale observations of 137Cs concentration. Results revealed that at the site, the Onsite wind field which captured the frequently changing wind, best reproduced the onsite gamma dose rates with the grid resolution of 200 m. At the local scale (up to 20 km), the observations present a smoother temporal change. The wind fields assimilated with Japanese domestic observations presented advantageous performance, and the 1-km NHM-LETKF achieved the best score of the factor of 5 metric of 0.49 for the simulated 137Cs concentration. The SPRAY coupled with the three-dimensional (3D) convolution method and RIMPUFF showed better performance in simulating the onsite gamma dose rate and the local-scale concentration, respectively. The ensemble mean achieved robust metrics, better simulated the baseline of onsite gamma dose rates, and reproduced more peaks of local-scale concentration at the expense of peak value deviation.
... Because of their spatiotemporal sparsity, ground-based observations alone might not capture all the transport processes. In this case, atmospheric transport modeling represents a key tool for analyzing radionuclide transport behavior and for predicting the environmental consequences (Korsakissok et al., 2013). ...
... For example, there are only a few Tokyo Electric Power Company (TEPCO) monitoring sites and one automated meteorological data acquisition system site within 20 km of the FDNPP, and 99.5% of the critical suspended particulate matter observations were acquired more than 20 km away from the FDNPP (Tsuruta et al., 2014). Direct local simulation of the area within 20 km of the FDNPP also confronts a series of technological challenges, e.g., reproduction of the frequently varying meteorological fields (Korsakissok et al., 2013) and modeling of the complex cross-scale transport behaviors at the site and up to 20 km range. ...
... The vertical extension of plume, on the other hand, is largely limited in the troposphere where convections are confined within. There are many literatures on horizontal transport, dispersion, deposition processes and uncertainty analysis [2][3][4][5], but only in a handful of model development papers can one find details describing modelling approaches on turbulent mixing [6][7][8][9] regarding the vertical extension of radioactive plume in the free troposphere. The vertical extension of plume, however, is closely related to the vertical mass distribution of pollutants and thus affects the concentration level in lower atmosphere as well as horizontal distribution. ...
Article
Dose estimation in the upper air is not studied as much as on ground level or in boundary layer. However, there is a need from stakeholders in aviation industry for a reasonable estimation of the radioactive plume impact at cruising levels. This study aims to provide a quantitative estimation of the dose and how reliable it is for dispersion processes up to seven days. A Lagrangian atmospheric dispersion model is used to estimate quantitively the vertical extension of radionuclides from simplified hypothetical radionuclide release scenarios. Sources at different latitudes are selected for simulation in a boreal winter case. Three meteorological data are examined to test the sensitivity of vertical plume distribution to driving meteorological data. The vertical distribution of air concentration of radionuclides is investigated and the associated uncertainties are analysed. It is found that the vertical extension of plumes is sensitive to meteorological data being used where vertical turbulent velocities play an important role. It is therefore necessary to address the uncertainties of air concentration or dose in the free troposphere and caution must be taken when providing the results to stakeholders.
... Many atmospheric dispersion models with different complexities have been used in representing atmospheric dispersive processes of radioactive materials, such as the Hybrid Single-Particle Lagrangian Integrated Trajectory (HYSPLIT) model, the FLEXible PARTicle dispersion model (FLEXPART), and the Gaussian puff model. They were used to analyze the atmospheric dispersion of radioactive materials for the Chernobyl and Fukushima NPP accidents [4][5][6][7]. In addition, the models have also been used to simulate radioactive dispersions of hypothetical accident scenarios for accidental preparedness and NPP site selection [8][9][10][11]. ...
Article
Background: High-fidelity meteorological data is a prerequisite for the realistic simulation of atmospheric dispersion of radioactive materials near nuclear power plants (NPPs). However, many meteorological models frequently overestimate near-surface wind speeds, failing to represent local meteorological conditions near NPPs. This study presents a new high-resolution (approximately 1 km) meteorological downscaling method for modeling short-range (< 100 km) atmospheric dispersion of accidental NPP plumes.Materials and Methods: Six considerations from literature reviews have been suggested for a new dynamic downscaling method. The dynamic downscaling method is developed based on the Weather Research and Forecasting (WRF) model version 3.6.1, applying high-resolution land-use and topography data. In addition, a new subgrid-scale topographic drag parameterization has been implemented for a realistic representation of the atmospheric surface-layer momentum transfer. Finally, a year-long simulation for the Kori and Wolsong NPPs, located in southeastern coastal areas, has been made for 2016 and evaluated against operational surface meteorological measurements and the NPPs’ on-site weather stations.Results and Discussion: The new dynamic downscaling method can represent multiscale atmospheric motions from the synoptic to the boundary-layer scales and produce three-dimensional local meteorological fields near the NPPs with a 1.2 km grid resolution. Comparing the year-long simulation against the measurements showed a salient improvement in simulating near-surface wind fields by reducing the root mean square error of approximately 1 m/s. Furthermore, the improved wind field simulation led to a better agreement in the Eulerian estimate of the local atmospheric dispersion. The new subgrid-scale topographic drag parameterization was essential for improved performance, suggesting the importance of the subgrid-scale momentum interactions in the atmospheric surface layer.Conclusion: A new dynamic downscaling method has been developed to produce high-resolution local meteorological fields around the Kori and Wolsong NPPs, which can be used in short-range atmospheric dispersion modeling near the NPPs.
Article
Radioactive Cs-rich microparticles (CsMPs) released from the Fukushima Daiichi Nuclear Power Plant (FDNPP) are a potential health risk through inhalation. Little has been documented on the occurrence of CsMPs, particularly their occurrence inside buildings. In this study, we quantitatively analyze the distribution and number of CsMPs in indoor dust samples collected from an elementary school located 2.8 km to the southwest of FDNPP. The school had remained deserted until 2016. Then, using a modified version of the autoradiography-based "quantifying CsMPs (mQCP) method," we collected samples and determined the number of CsMPs and Cs radioactive fraction (RF) values of the microparticles (defined as total Cs activity from CsMPs/bulk Cs activity of the entire sample). The numbers of CsMPs ranged from 653 to 2570 particles/(g dust) and 296-1273 particles/(g dust) on the first and second floors of the school, respectively. The corresponding RFs ranged between 6.85 - 38.9% and 4.48-6.61%, respectively. The number of CsMPs and RF values in additional outdoor samples collected near the school building were 23-63 particles/(g dust or soil) and 1.14-1.61%, respectively. The CsMPs were most abundant on the school's first floor near to the entrance, and the relative abundance was higher near the stairs on the second floor, indicating a likely CsMP dispersion path through the building. Additional wetting of the indoor samples combined with autoradiography revealed that indoor dusts had a distinct absence of intrinsic, soluble Cs species, such as CsOH. These combined observations indicate that a significant amount of poorly soluble CsMPs were likely contained in initial radioactive airmass plumes from the FDNPP and that the microparticles penetrated buildings. CsMPs could still be abundant at the location, with locally high Cs activity in indoor environments near to openings.
Article
Full-text available
We report a detailed dynamics study on the mode-specificity of the HI + C2H5 two-channel reaction (H-abstraction and I-abstraction), through performing quasi-classical trajectory computations on a recently developed high-level ab initio full-dimensional spin-orbit-corrected potential energy surface, by exciting four different vibrational modes of reactants at five collision energies. The effect of the normal-mode excitations on the reactivity, the mechanism, and the post-reaction energy flow is investigated. Both reaction pathways are intensely promoted when the HI-stretching mode is excited while the excitations imposed on C2H5 somewhat surprisingly inhibit the dominant H-abstraction reaction pathway. The enhancement effect of the excitation in the HI vibrational mode is found to be much more effective than increasing the translational energy, similar to the HBr + C2H5 reaction. Not like the Br-abstraction pathway, however, the I-abstraction reaction pathway could be comparable to the dominant H-abstraction reaction pathway. The dominance of the direct stripping mechanism is indicated in H-abstraction while the direct rebound mechanism is observed in I-abstraction. The H-abstraction is much pickier about the initial attack angle distributions for HI than I-abstraction is, which leads to a decrease in reactivity in the H-abstraction reaction pathway. The dominance of side-on CH3CH2 attack in I-abstraction is more obvious than in the case of H-abstraction. In the case of the H-abstraction reaction pathway, the major part of the initial translational energy ends up in translational recoil, while for I-abstraction most energy excites the product C2H5I.
Article
For the purpose of realizing atmospheric radiation plume tracking, laser-induced breakdown spectroscopy (LIBS) was performed to characterize Cs atomic emissions from airborne nanoparticles in a binary particle matrix using nanosecond and femtosecond laser ablation. As a surrogate to multi-element nuclear fallout particulates, both Cs and Na solutions were mixed with Cu at a 1:1.45 M concentration ratio and aerosolized to study elemental fractionation in atomic emissions, revealing Na I resonance emissions are enhanced whereas Cs I resonance emissions are reduced with Cu present despite sharing similar electronic and chemical properties. Femtosecond laser ablation reduced elemental fractionation effects by more than 50% while retaining the same temporal fractionation trends for Cs and Na. The origin of elemental fractionation in aerosol LIBS was further investigated using Al and Ti aerosols which revealed that the particle matrix induces changes to the local plasma temperature, population of atomic states, and subsequent molecular association and emission.
Article
Full-text available
On March 11, 2011, an earthquake and tsunami hit the northeast coast of Japan and damaged the Fukushima Daiichi nuclear power plant, leading to the release of radioactive material into the atmosphere. We trace the evolution of radioactivity release to the atmosphere and subsequent dispersion as simulated by models, and we compare these to actual measurements. Four main release periods are highlighted. The first event had limited consequences to the north of the power plant along the coast; the second had no impact on Japanese territory because the plumes travelled toward the Pacific Ocean; the third was responsible for significant and long-term impact, especially northwest of the plant; and the last had consequences of lesser impact on the Tokyo area.
Article
Full-text available
A major difficulty when inverting the source term of an atmospheric tracer dispersion problem is the estimation of the prior errors: those of the atmospheric transport model, those ascribed to the representativity of the measurements, those that are instrumental, and those attached to the prior knowledge on the variables one seeks to retrieve. In the case of an accidental release of pollutant, the reconstructed source is sensitive to these assumptions. This sensitivity makes the quality of the retrieval dependent on the methods used to model and estimate the prior errors of the inverse modeling scheme. We propose to use an estimation method for the errors' amplitude based on the maximum likelihood principle. Under semi-Gaussian assumptions, it takes into account, without approximation, the positivity assumption on the source. We apply the method to the estimation of the Fukushima Daiichi source term using activity concentrations in the air. The results are compared to an L-curve estimation technique and to Desroziers's scheme. The total reconstructed activities significantly depend on the chosen method. Because of the poor observability of the Fukushima Daiichi emissions, these methods provide lower bounds for cesium-137 and iodine-131 reconstructed activities. These lower bound estimates, 1.2 × 1016 Bq for cesium-137, with an estimated standard deviation range of 15%-20%, and 1.9 - 3.8 × 1017 Bq for iodine-131, with an estimated standard deviation range of 5%-10%, are of the same order of magnitude as those provided by the Japanese Nuclear and Industrial Safety Agency and about 5 to 10 times less than the Chernobyl atmospheric releases.
Article
Full-text available
On 11 March 2011, an earthquake occurred about 130 km off the Pacific coast of Japan's main island Honshu, followed by a large tsunami. The resulting loss of electric power at the Fukushima Dai-ichi nuclear power plant (FD-NPP) developed into a disaster causing massive release of radioactivity into the atmosphere. In this study, we determine the emissions of two isotopes, the noble gas xenon-133 (133Xe) and the aerosol-bound caesium-137 (137Cs), which have very different release characteristics as well as behavior in the atmosphere. To determine radionuclide emissions as a function of height and time until 20 April, we made a first guess of release rates based on fuel inventories and documented accident events at the site. This first guess was subsequently improved by inverse modeling, which combined the first guess with the results of an atmospheric transport model, FLEXPART, and measurement data from several dozen stations in Japan, North America and other regions. We used both atmospheric activity concentration measurements as well as, for 137Cs, measurements of bulk deposition. Regarding 133Xe, we find a total release of 16.7 (uncertainty range 13.4-20.0) EBq, which is the largest radioactive noble gas release in history not associated with nuclear bomb testing. There is strong evidence that the first strong 133Xe release started very early, possibly immediately after the earthquake and the emergency shutdown on 11 March at 06:00 UTC. The entire noble gas inventory of reactor units 1-3 was set free into the atmosphere between 11 and 15 March 2011. For 137Cs, the inversion results give a total emission of 35.8 (23.3-50.1) PBq, or about 42% of the estimated Chernobyl emission. Our results indicate that 137Cs emissions peaked on 14-15 March but were generally high from 12 until 19 March, when they suddenly dropped by orders of magnitude exactly when spraying of water on the spent-fuel pool of unit 4 started. This indicates that emissions were not only coming from the damaged reactor cores, but also from the spent-fuel pool of unit 4 and confirms that the spraying was an effective countermeasure. We also explore the main dispersion and deposition patterns of the radioactive cloud, both regionally for Japan as well as for the entire Northern Hemisphere. While at first sight it seemed fortunate that westerly winds prevailed most of the time during the accident, a different picture emerges from our detailed analysis. Exactly during and following the period of the strongest 137Cs emissions on 14 and 15 March as well as after another period with strong emissions on 19 March, the radioactive plume was advected over Eastern Honshu Island, where precipitation deposited a large fraction of 137Cs on land surfaces. The plume was also dispersed quickly over the entire Northern Hemisphere, first reaching North America on 15 March and Europe on 22 March. In general, simulated and observed concentrations of 133Xe and 137Cs both at Japanese as well as at remote sites were in good quantitative agreement with each other. Altogether, we estimate that 6.4 TBq of 137Cs, or 19% of the total fallout until 20 April, were deposited over Japanese land areas, while most of the rest fell over the North Pacific Ocean. Only 0.7 TBq, or 2% of the total fallout were deposited on land areas other than Japan.
Article
Full-text available
On 11 March 2011, an earthquake occurred about 130 km off the Pacific coast of Japan's main island Honshu, followed by a large tsunami. The resulting loss of electric power at the Fukushima Dai-ichi nuclear power plant (FD-NPP) developed into a disaster causing massive release of radioactivity into the atmosphere. In this study, we determine the emissions of two isotopes, the noble gas xenon-133 (133Xe) and the aerosol-bound caesium-137 (137Cs), which have very different release characteristics as well as behavior in the atmosphere. To determine radionuclide emissions as a function of height and time until 20 April, we made a first guess of release rates based on fuel inventories and documented accident events at the site. This first guess was subsequently improved by inverse modeling, which combined the first guess with the results of an atmospheric transport model, FLEXPART, and measurement data from several dozen stations in Japan, North America and other regions. We used both atmospheric activity concentration measurements as well as, for 137Cs, measurements of bulk deposition. Regarding 133Xe, we find a total release of 16.7 (uncertainty range 13.4-20.0) EBq, which is the largest radioactive noble gas release in history not associated with nuclear bomb testing. There is strong evidence that the first strong 133Xe release started very early, possibly immediately after the earthquake and the emergency shutdown on 11 March at 06:00 UTC. The entire noble gas inventory of reactor units 1-3 was set free into the atmosphere between 11 and 15 March 2011. For 137Cs, the inversion results give a total emission of 35.8 (23.3-50.1) PBq, or about 42% of the estimated Chernobyl emission. Our results indicate that 137Cs emissions peaked on 14-15 March but were generally high from 12 until 19 March, when they suddenly dropped by orders of magnitude exactly when spraying of water on the spent-fuel pool of unit 4 started. This indicates that emissions were not only coming from the damaged reactor cores, but also from the spent-fuel pool of unit 4 and confirms that the spraying was an effective countermeasure. We also explore the main dispersion and deposition patterns of the radioactive cloud, both regionally for Japan as well as for the entire Northern Hemisphere. While at first sight it seemed fortunate that westerly winds prevailed most of the time during the accident, a different picture emerges from our detailed analysis. Exactly during and following the period of the strongest 137Cs emissions on 14 and 15 March as well as after another period with strong emissions on 19 March, the radioactive plume was advected over Eastern Honshu Island, where precipitation deposited a large fraction of 137Cs on land surfaces. The plume was also dispersed quickly over the entire Northern Hemisphere, first reaching North America on 15 March and Europe on 22 March. In general, simulated and observed concentrations of 133Xe and 137Cs both at Japanese as well as at remote sites were in good quantitative agreement with each other. Altogether, we estimate that 6.4 TBq of 137Cs, or 19% of the total fallout until 20 April, were deposited over Japanese land areas, while most of the rest fell over the North Pacific Ocean. Only 0.7 TBq, or 2% of the total fallout were deposited on land areas other than Japan.
Article
The 'source term' including the time evolution of the release rate to the atmosphere and its distribution between radioisotopes remains one of the key uncertainties in the understanding of the consequences of the Fukushima Dai-ichi accident. Inverse modeling methods have already proved to be efficient to estimate accidental releases. This paper presents a new inverse modeling approach to assess the source term by using gamma dose rate monitoring. The approach is applied to the Fukushima accident. The reliability of the inverted source term is estimated by using model to data comparison and yields a good agreement. An important outcome on this study is its applicability during a response to an emergency situation.
Article
It has been postulated that atmospheric pathways may comprise a significant source of nitrogen for aquatic ecosystems and excess atmospheric deposition to coastal areas may be a major cause of eutrophication. Dry deposition of nitrogen containing particles is a potential, but poorly quantified pathway, for atmospheric nitrogen flux. This pathway is not well quantified because deposition velocities for particles are difficult to calculate and incorporate substantial uncertainties. Herein we employ an amended version of the Hummelshøj et al. (1992, Proceedings of the 5th International Conference on Precipitation Scavenging and Atmosphere–Surface Exchange Processes. AMS, Richland, Washington, USA, 12pp.) model to calculate size-segregated dry deposition of particle inorganic nitrogen compounds to the western Baltic during the late Spring of 1997 based on data collected as part of the Air–Sea Exchange Process Study (ASEPS). The results show that over a 15d period in April and May dry deposition fluxes varied between 30 and 400μgm-2d-1 for nitrate and 1 and 120μgm-2d-1 for ammonium. Sensitivity analyses run to assess the potential bounds on actual dry deposition indicate that, for reasonable variation of model parameters and formulation, particle nitrogen dry deposition may be varied by up to an order of magnitude. The primary sources of uncertainty are identified and are discussed in the context of alternative model formulations.
Article
To understand the atmospheric behavior of radioactive materials emitted from the Fukushima Daiichi nuclear power plant after the nuclear accident that accompanied the great Tohoku earthquake and tsunami on 11 March 2011, we simulated the transport and deposition of iodine-131 and cesium-137 using a chemical transport model. The model roughly reproduced the observed temporal and spatial variations of deposition rates over 15 Japanese prefectures (60-400 km from the plant), including Tokyo, although there were some discrepancies between the simulated and observed rates. These discrepancies were likely due to uncertainties in the simulation of emission, transport, and deposition processes in the model. A budget analysis indicated that approximately 13% of iodine-131 and 22% of cesium-137 were deposited over land in Japan, and the rest was deposited over the ocean or transported out of the model domain (700 × 700 km2). Radioactivity budgets are sensitive to temporal emission patterns. Accurate estimation of emissions to the air is important for estimation of the atmospheric behavior of radionuclides and their subsequent behavior in land water, soil, vegetation, and the ocean.