Sunteți pe pagina 1din 15

Journal of Environmental Radioactivity 136 (2014) 41e55

Contents lists available at ScienceDirect

Journal of Environmental Radioactivity


journal homepage: www.elsevier.com/locate/jenvrad

Modeling the fallout from stabilized nuclear clouds using the HYSPLIT
atmospheric dispersion model
G.D. Rolph a, *, F. Ngan a, b, R.R. Draxler a
a
NOAA Air Resources Laboratory (ARL), NCWCP R/ARL, 5830 University Research Court, College Park, MD 20740, USA
b
Cooperative Institute for Climate and Satellites, University of Maryland, College Park, MD, USA

a r t i c l e i n f o a b s t r a c t

Article history: The Hybrid Single Particle Lagrangian Integrated Trajectory (HYSPLIT) model, developed by the National
Received 21 February 2014 Oceanic and Atmospheric Administrations Air Resources Laboratory, has been congured to simulate the
Received in revised form dispersion and deposition of nuclear materials from a surface-based nuclear detonation using publicly
6 May 2014
available information on nuclear explosions. Much of the information was obtained from The Effects of
Accepted 7 May 2014
Available online 28 May 2014
Nuclear Weapons by Glasstone and Dolan (1977). The model was evaluated against the measurements
of nuclear fallout from six nuclear tests conducted between 1951 and 1957 at the Nevada Test Site using
the global NCEP/NCAR Reanalysis Project (NNRP) and the Weather Research and Forecasting (WRF)
Keywords:
HYSPLIT
meteorological data as input. The model was able to reproduce the general direction and deposition
Dispersion modeling patterns using the coarse NNRP data with Figure of Merit in Space (FMS e the percent overlap between
Nuclear fallout predicted and measured deposition patterns) scores in excess of 50% for four of six simulations for the
Deposition smallest dose rate contour, with FMS scores declining for higher dose rate contours. When WRF mete-
orological data were used the FMS scores were 5e20% higher in ve of the six simulations, especially at
the higher dose rate contours. The one WRF simulation where the scores declined slightly (10e30%) was
also the best scoring simulation when using the NNRP data. When compared with measurements of dose
rate and time of arrival from the Town Data Base (Thompson et al., 1994), similar results were found with
the WRF simulations providing better results for four of six simulations. The overall result was that the
different plume simulations using WRF data had more consistent performance than the plume simu-
lations using NNRP data elds.
Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/3.0/).

1. Introduction radioactive debris measured by Air Weather Service B-29 weather


reconnaissance aircraft. As this collaboration progressed, it became
The National Oceanic and Atmospheric Administrations (NOAA) clear that the development of more complex atmospheric transport
Air Resources Laboratory (ARL) has its origins in providing meteo- models was needed to predict the long-range transport and
rological support to other Federal agencies during the Cold Wars deposition of nuclear materials from the many tests being con-
nuclear arms race. In 1948, a predecessor of ARL, the U.S. Weather ducted around the world, and over the next several decades ARL
Bureau created a Special Projects Section (SPS), in Washington, D.C., maintained a continued interest in the transport, dispersion, and
to bridge the gap between the meteorological expertise of the deposition of nuclear materials (Draxler, 1982; Ferber and Heffter,
Weather Bureau and the research needs of other agencies related to 1961, 1976; Heffter, 1969, 1980; Hoecker and Machta, 1990;
nuclear weapons testing and development. As the arms race Machta and List, 1956; Machta et al., 1962; Machta and Heffter,
accelerated, a need arose to determine where clandestine nuclear 1986; Telegadas et al., 1978, 1979; Telegadas and List, 1964; List
tests were conducted using measurements of nuclear fallout et al., 1961).
around the world. The SPS was involved in determining the location With an increase in the concern for future terrorist incidents
of the rst Soviet nuclear test (Machta, 1992) conducted in 1949 since the September 11, 2001, attack on the World Trade Center in
using backward trajectory analysis from interceptions of New York City and the Pentagon in Washington, D.C., and the
possibility that terrorists could acquire nuclear materials and
develop an Improvised Nuclear Device (IND), ARL expanded its
* Corresponding author. Tel.: 1 301 683 1376. capabilities for responding to exercises and incidents involving the
E-mail address: glenn.rolph@noaa.gov (G.D. Rolph). detonation of a nuclear device and made them available to the

http://dx.doi.org/10.1016/j.jenvrad.2014.05.006
0265-931X/Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/).
42 G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55

civilian forecasters of the U.S. National Weather Service in support as the operational dispersion model for the National Weather
of their local clients. Service providing routine forecasts of smoke from wild res (Rolph
A number of models have been used to simulate the explosion et al., 2009; Stein et al., 2009), windblown dust (Draxler et al.,
and fallout from a nuclear weapon with varying degrees of 2010), volcanic ash (Stunder et al., 2007), and atmospheric
complexity. For example, the HotSpot Gaussian plume model dispersion products for chemical and nuclear accidents (Draxler
(Homann and Aluzzi, 2013) developed by the U.S. Department of and Rolph, 2012).
Energy is designed to provide a quick response to eld personnel on The focus of the research presented here will be to model the
the radiation effects from the atmospheric release of radioactive dispersion, deposition, and decay of nuclear debris that followed
materials and is designed for short-range (less than 10 km), and the detonation of six relatively small (<50 kT) nuclear devices in
short-term (less than a few hours) predictions. A special purpose the 1950s in Nevada using the HYSPLIT model and to calculate the
program is included in HotSpot to model the effects of a surface- radioactive dose rates from the local fallout using a new module
burst nuclear weapon; however the model assumes a constant developed for calculating doses in HYSPLIT. The resulting dose rate
wind direction and speed and does not take into account the effects patterns will then be compared to the measured dose rate patterns
of terrain on the wind ow. from each of the six Nevada nuclear tests. The goal is to parame-
A much more sophisticated model is the Hazard Prediction & terize the HYSPLIT model using information currently available in
Assessment Capability (HPAC) model developed by the Defense the literature about the nuclear source term so that it can be run
Threat Reduction Agency (DTRA, 2001; Chang et al., 2005). HPAC, quickly to give a realistic estimate of the magnitude and pattern of
which is a composed of a hazard source denition module, a deposition from small nuclear devices. Therefore, HYSPLIT will not
transport module, and an effects module, uses the Second-order be used to model the initial nuclear cloud growth, but will assume
Closure Integrated PUFF (SCIPUFF) dispersion model to quickly that the nuclear cloud has stabilized before proceeding with the
predict the hazards from nuclear, biological, chemical, and radio- transport and deposition calculations. In addition, the model will
logical weapons and facilities. The second-order turbulence closure be run with global- and local-scale gridded meteorological data to
theory employed by SCIPUIFF provides a unique advantage over determine if increasing the horizontal and vertical resolution of the
other models in that it can also produce a probabilistic estimate of meteorological data will result in improved predictions of nuclear
the uncertainty in the concentration results due to atmospheric fallout.
dispersion. HPAC can input real-time observed wind speed and
direction at a single location or use 3-dimensional gridded wind 2. Dispersion model experimental design
and temperature elds to calculate the pollutant transport. HPAC
uses the Nuclear Weapon (NWPN) module to calculate the initial 2.1. HYSPLIT model
stabilized nuclear cloud prior to modeling the dispersion and
deposition by SCIPUFF. The Hybrid Single-Particle Lagrangian Integrated Trajectory
Recently, the Norwegian Meteorological Institute (Bartnicki and (HYSPLIT) model is a Lagrangian particle and puff model that is
Saltbones, 2008) modied the Severe Nuclear Accident Program used by air quality researchers and forecasters to model the
(SNAP) model to not only predict the dispersion of radioactive transport of pollutants using 3-dimensional gridded meteorolog-
debris following a nuclear accident, but also from a nuclear ex- ical elds. In a recent publication (Moroz et al., 2010), the model
plosion. The model assumes two types of stabilized cloud shapes; a was used to reconstruct the 137Cs deposition from nuclear tests in
cylinder and a mushroom shape. Particle sizes range from 2 to the Marshall Islands, the Nevada Test Site, and the Semipalatinsk
200 mm in ten discrete size ranges and an activity distribution that Test Site of the former Soviet Union. HYSPLIT was congured with a
assumes 10% of the activity is in each category. They found that the log-normal particle distribution and each particle size bin was
results were not sensitive to different initial nuclear cloud shapes, assigned a fraction of the total 137Cs activity. The resulting 137Cs
however there were considerable differences in the plumes for modeled deposition patterns were then compared to the observed
various nuclear yields due to transport at different levels in the patterns from each nuclear test. The study suggested that given
atmosphere. adequate spatial and temporal meteorological data, HYSPLIT can be
Finally, one of the most advanced nuclear fallout models is the used to determine where and when nuclear material was deposited
Defense Land Fallout Interpretive Code (DELFIC, Norment, 1979a, with relatively good degree of accuracy. They also found that when
1979b), which was originally developed in 1968 by the Defense no measurements are available, HYSPLIT can be used to determine
Atomic Support Agency and has been subsequently revised by the if fallout might have occurred at a given location and also some
military, government laboratories, and private organizations. indication of the magnitude of the deposition.
DELFIC includes a dynamic, one-dimensional cloud rise module In this application, we expand on the work of Moroz et al. (2010)
that explicitly calculates the growth of the nuclear cloud based on and congure the model with several particle size and activity
the yield of the weapon, the soil type, the height of burst, the at- distributions obtained from various published sources, and
mospheric prole, the particle size distribution, etc. DELFIC is compute dose rate contours for several nuclear tests at the Nevada
intended for use in the research of local nuclear fallout prediction Test Site in Nevada, USA.
and to be the standard against which simpler models are compared.
DELFIC computes the downwind spread of nuclear material based 2.2. Stabilized nuclear cloud
on the assumed particle fall velocities and turbulence in the
downwind direction from the calculated cloud rise module. To model the nuclear fallout quickly it is necessary to establish
One of the atmospheric dispersion models developed by ARL in the characteristics of the nuclear cloud just after stabilization
the late 1970s, and the primary model in use by NOAA today, is the instead of calculating the complex evolution of the cloud beginning
Hybrid Single Particle Lagrangian Integrated Trajectory (HYSPLIT) at the time of detonation. Stabilization occurs after the temperature
(Draxler, 1999; Draxler and Hess, 1997, 1998) model. This model of the nuclear cloud is equalized with the ambient temperature in
originally used surface weather observations and upper-air the surrounding air. At this point entrainment of outside air ceases
soundings as its source of meteorology until gridded meteorolog- and the vertical growth of the cloud stops. Nuclear particles have
ical model forecast data became routinely available from the U.S. been formed directly by the ssion reaction, the condensation of
National Weather Service in the late 1980s. Today, HYSPLIT serves vaporized material lofted from the surface, and the vaporized
G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55 43

components of the device itself. These particles settle out of the initial particle size. For example, in this study its value is set to 5,
atmosphere at the particles fall velocity, which is based on the size, therefore for each of the initial 14 particles, 5 particles are created
shape, and density of the particle. In the HYSPLIT calculations, all with sizes centered on the initial particle diameter and the mass of
particles are assumed to be spherical and have a density of 2.5 g/ each particle is dened by assuming that dV/d(log R) is linear be-
cm3 (Glasstone and Dolan, 1977) assuming that the bulk of the tween the dened points for an increasing cumulative mass dis-
radioactive material is attached to soil particles. tribution with respect to particle diameter. As a result of this
The normalized nuclear activity distribution in the typical calculation, HYSPLIT denes seventy particle size bins from the
mushroom-shaped nuclear cloud is adapted from the ARL Fallout initially dened fourteen.
Prediction Technique (AFPT: Heffter, 1969). AFPT denes the initial In Table 2, initial log-mean particle diameters are shown for the
nuclear cloud as a cylinder that is divided into six layers (disks) default particle distribution obtained from Fig. 9.164 of Glasstone
between the surface and the top of the stabilized nuclear cloud and Dolan (1977). As mentioned earlier, dry particle removal is
(Fig. 1). The top and bottom of the cap of the cloud are dened calculated through a particle gravitational settling velocity based
based on the NATO (2010) operations manual. The thickness and on the particle radius, density (2.5 g/cm3), and shape (spherical).
height of each layer depends on the nuclear yield and are deter- Although wet removal of particles is not considered in this study
mined by linearly interpolating between the top and bottom of the (the Nevada tests were designed to be conducted when no rain was
cap and between the surface and the bottom of the cap. For forecast near ground zero within 10 h of detonation, Fehner and
example, for a 10 kT explosion (Table 1), the layers range from Gosling, 2006), wet removal can be dened in HYSPLIT using an
1700 m thick near the surface to 1000 m thick at the top of the cap in-cloud rainout ratio (ratio of pollutant in air to that in rain
(8200 m m above ground level). measured at the ground) and a below-cloud washout removal time
One simplication that was assumed for these simulations is constant. Typical values for 137Cs are 4.0  104 l/l for in-cloud and
that all particles are released from a vertical line through the center 5.0  106 s1 for below-cloud removal (Draxler and Rolph, 2012).
of the cloud instead of over the entire diameter of the cloud. This
assumption should not impact the results substantially as the
concentration grid size (w8 km) used to display the results is larger 2.4. Activity distribution
than the width of the cloud (<3 km, Glasstone and Dolan, 1977) for
the size of the yields being studied here. In addition to a particle size distribution, it is necessary to dene
an activity distribution within the stabilized nuclear cloud. Each
particle size bin in each vertical cloud layer is assigned an initial
2.3. Particle distribution and removal activity. Fig. 1 shows the activity in each cloud layer as a percentage
of the total activity in the cloud as dened by Heffter (1969). The
Within each cloud layer, HYSPLIT initially denes fourteen types majority (78%) of the activity is initially distributed in the top half of
of particles or parcels to represent the radionuclide distribution: the nuclear cloud. HYSPLIT can be congured with any user dened
one particle represents all noble gases (no settling velocity) and the particle activity distribution, however, for this study only ve of the
other thirteen represent various particle size bins depending on the most common particle activity distributions (Fig. 2) were tested
choice of the predened particle distribution. Although fourteen with HYSPLIT: Glasstone and Dolan (1977); Izrael (2002); Miller
particle size bins could be used to run the nuclear simulations, and Sartor (1965); Heffter (1969), and DELFIC (Jodoin, 1994;
sensitivity tests revealed that using more particle bins provides a Norment, 1979a, 1979b). The Glasstone, Izrael, and DELFIC activity
more continuous representation of the mass distribution in the distributions are log-normal, whereas the Small Boy and Heffter
vertical. A new HYSPLIT option that was developed for this appli- activity distributions were derived from fallout information from
cation allows the user to redistribute each particle size bin into the Small Boy shot in July, 1962, and are not log-normal. As an
several new particle size bins centered about the initial particle example, Table 2 shows the activity distribution for the default
size. In addition, the mass dened on each initial particle is linearly Glasstone distribution for each particle size and cloud layer. Most of
distributed amongst the newly dened particles. A variable is set by the activity is in the upper layers (>6000 m) and the small- to mid-
the user to dene the number of new size bins assigned to each sized particles (<100 mm).
It should also be noted that the model assumes the same
composition of radionuclides exists on each particle, regardless of
the particle size. In reality, as has been discussed by Hicks (1981,
1984) and Trabalka and Kocher (2007), fractionation occurs as the
vaporized soil material condenses at varying rates with the re-
fractory radionuclides, which have high boiling points, forming
solid particles onto which the volatile radionuclides with lower
boiling points condense. Fractionation causes larger particles con-
taining a higher fraction of refractory radionuclides to fallout closer
to ground zero while the smaller particles containing the more
volatile radionuclides continue farther downwind. Although other
models (Moroz et al., 2010) have accounted for fractionation by
removing a fraction of the activity of volatile radionuclides on
larger particles and enhancing the activity on smaller particles, it
was not accounted for in this initial model development due to the
uncertainty in the amount removed due to this complex process.
A review of the 213 nuclear species present in the nuclear cloud
one minute after detonation (England and Rider, 1993) indicated
Fig. 1. Model representation of a nuclear mushroom cloud and the total activity
that approximately 83% of the nuclear activity is present as Noble
fraction in each vertical segment dened by the model for a 10 kT explosion (Heffter, gases. Therefore the activity distribution numbers given in Table 2
1969). The shaded region represents the layers dening the cap of the cloud. were corrected to reect this mix by dening the generic Noble gas
44 G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55

Table 1
Model cloud layer heights (meters above ground level) based on nuclear yield in kilotons (kT).

Yield (kT) 2.5 7.5 12.5 17.5 22.5 27.5 32.5 37.5 42.5 45.0

Level 7 3700 6300 8200 9700 10800 11200 11600 11900 12200 12500
Level 6 3132 5434 7166 8532 9532 9900 10232 10500 10766 11000
Level 5 2566 4567 6130 7366 8266 8600 8866 9100 9333 9500
Level 4 2000 3700 5100 6200 7000 7300 7500 7700 7900 8000
Level 3 1334 2466 3400 4132 4666 4866 5000 5132 5266 5332
Level 2 667 1233 1700 2066 2333 2433 2500 2566 2633 2666
Level 1 0 0 0 0 0 0 0 0 0 0

Table 2
Initial log-mean particle diameter (mm) and percentage of nuclear activity by cloud layer and particle size for the Glasstone particle distribution. The total activity sums to 100%.

Diameter (mm) 500 350 275 225 187.5 162.5 137.5 112.5 87.5 70 57.5 45 20

Height (m) P1 P2 P3 P4 P5 P6 P7 P8 P9 P10 P11 P12 P13 Total

8200 0.175 0.525 0.525 0.875 0.700 1.050 1.400 2.100 3.150 1.750 1.750 1.400 2.100 17.5
7166 0.300 0.900 0.900 1.500 1.200 1.800 2.400 3.600 5.400 3.000 3.000 2.400 3.600 30
6130 0.300 0.900 0.900 1.500 1.200 1.800 2.400 3.600 5.400 3.000 3.000 2.400 3.600 30
5100 0.150 0.450 0.450 0.750 0.600 0.900 1.200 1.800 2.700 1.500 1.500 1.200 1.800 15
3400 0.050 0.150 0.150 0.250 0.200 0.300 0.400 0.600 0.900 0.500 0.500 0.400 0.600 5
1700 0.250 0.075 0.075 0.125 0.100 0.150 0.200 0.300 0.450 0.250 0.250 0.200 0.300 2.5
Total 1 3 3 5 4 6 8 12 18 10 10 8 12 100

particle (Pgas) with 83% of the activity and proportionally (Table 3), which are used to compute various doses at a later step.
reducing the other particle activity values. This allows the user to modify the dose calculations, say with other
HYSPLIT uses an emissions text le to dene the initial activity dose conversion factors or breathing rates, without having to rerun
distribution for each particle size in each layer in units of activity the model.
per hour. The model assumes a 1 min emission for the nuclear
explosion simulation (the initial emission values in activity per
minute are multiplied by 60 to obtain activity per hour). 2.5. Fission yield
The default HYSPLIT nuclear explosion simulation is initially
congured with 9000 particles distributed over the 70 particle size Prior to being able to compute doses, it is necessary to compute
bins. However, after HYSPLIT calculates the number of particles the radionuclide yield for various ssion reactions, a step that only
needed to represent the source term through the six vertical layers, needs to be performed once. The calculation is based upon the yield
approximately 54,000 three-dimensional particles are required. In per 100 ssions as taken from England and Rider (1993), which has
addition, the HYSPLIT nuclear simulation is performed assuming a been digitally tabulated and available from http://ie.lbl.gov/ssion.
unit source emission; the yield of the weapon is not dened prior to html. Radionuclide inventories were computed using the rate of
computing the transport and deposition except in terms of dening 1.45  1023 ssions per kT (Glasstone and Dolan, 1977). The result is
the top of the cloud (Table 1). The model produces the dilution and a table (not shown) of the activity for a one kT detonation in Bec-
deposition factors between the ground and 100 m AGL at various querels (Bq) for 213 radionuclides with one column each for 235U
times downwind in terms of a unit emission on eight grids and 239Pu for high-energy (blast) and thermal (reactor) ssions,
respectively. Other columns have been added to the table for
radionuclide half-lives and the external dose rate conversion fac-
tors for cloud- and ground-shine as obtained from Eckerman and
Leggett (1996). Although most nuclear weapons test were not
pure 235U and had some of their beta activity from activation
products (Hicks, 1981; Glasstone and Dolan, 1977), for this simu-
lation, only the high-energy 235U activities were considered.

Table 3
Concentration (dilution) and deposition model dened grid details.

Type Grid resolution Grid area Interval Duration


(deg. latitude) (deg. latitude) (hr) (hr)

GRID1 Dilution 0.02 10  10 1 12


GRID2 Dilution 0.08 30  60 1 24
GRID3 Deposition 0.02 10  10 1 12
GRID4 Deposition 0.08 30  60 1 24
GRID5 Dilution & 0.02 10  10 1 12
Deposition
GRID6 Dilution & 0.08 30  60 1 24
Deposition
GRID7 Deposition 0.02 10  10 24 24
Fig. 2. Particle activity distributions used by the model: Glasstone, Izrael, Small Boy,
GRID8 Deposition 0.08 30  60 24 24
Heffter, and DELFIC.
G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55 45

2.6. Dose rate calculations was the only surface shot detonated at the NTS and only had a yield
of 1.2 kT, although it did produce signicant doses immediately
As previously mentioned, the resulting HYSPLIT output les downwind of ground zero. In the following sections, details on each
contain the dilution and deposition factors at various times down- nuclear test used in this study are presented, including information
wind in terms of a unit source emission. In order to produce a dose or on the yield of the weapon, the weather conditions, and the
dose rate, a post-processing program converts these elds to a dose observed fallout patterns. For more information on each shot, see
by multiplying the actual yield (in kT) by the ssion reaction (235U or the references provided in each section.
239
Pu) activity. The activity for each radionuclide is decayed from the
detonation time to each concentration averaging period using the 3.1. Buster-Jangle Sugar shot
half-life of each species dened in the activity le. The species
dependent dose rate is then computed using the Eckerman and The rst nuclear weapons test to be detonated at the surface
Leggett (1996) dose conversion factors from the air concentration (1 m) was the Buster-Jangle Sugar series shot on November 19, 1951.
(for cloud-shine & inhalation doses) and deposition (for ground- This device, detonated at 1700 UTC, had a yield of 1.2 kT, the cloud
shine dose). The cloud- and ground-shine dose conversion factors reached to a height of 4572 m MSL, and created a crater 6 m deep
do not include any contribution from the ingrowth of daughter and 27 m across. Given the small yield and detonation height near
products in the environment, although the inhalation dose con- the ground, this shot was considered a good candidate for this
version factors do account for the contribution from ingrowth in the study.
body after inhalation. The nal dose is integrated over all species. A The observed surface winds at detonation time (t0) were 1 m s1
time delay may also be specied to increase the decay time, allowing from the south, 16 m s1 from the south-southwest at w3000 m
the user to compare results with measurements that may apply to a MSL, and 21 m s1 from the south-southwest at w4250 m MSL
particular time after detonation. The user has the option to display (Ponton et al., 1982a; Hawthorne, 1979). Extremely high levels of
the following maps: particle animation; Total Effective Dose radioactivity (7500 R per hour, R h1) were initially observed just
Equivalent (TEDE); 4-day dose from deposition; cloud-shine due to north of ground zero (Fehner and Gosling, 2006), but fell rapidly
immersion in the plume, inhalation & ground-shine doses or dose after detonation. Unfortunately, there were very few radioactivity
rates; 24 h dose from deposition; dose rate from ground-shine after measurements off-site, so the off-site pattern is less reliable;
6 h; and decayed concentration & deposition. however measured radiation intensities ranged from 1 R h1 to
0.01 R h1 up to 6.3 km north of ground zero.
2.7. Meteorological input data

The availability of three-dimensional gridded meteorological 3.2. Tumbler-Snapper Easy shot


data to run HYSPLIT is limited during the 1950s and 1960s. For this
study, the model was initially run using meteorological data from the The next weapons test simulated here was the Operation
global NCEP/NCAR Reanalysis Project (NNRP; Kalnay et al., 1996). The Tumbler-Snapper Easy shot on May 7, 1952, at 1215 UTC. This
NNRP data were created from a joint venture between the NOAA weapon, which had a yield of 12 kT and extended to 10 km MSL,
National Centers for Environmental Prediction (NCEP) and the Na- was detonated from a 91 m tower. Because of strong upper-level
tional Center for Atmospheric Research (NCAR) with the goal of winds from the southwest, higher than expected levels of
producing an archive of meteorological data from 1948 onwards radioactive fallout (800 mR h1) were observed over the small
using a consistent suite of data assimilation and atmospheric models. community of Lincoln Mine, Nevada, 72 km to the northeast of
The NNRP data, originally in GRIB format, have been converted ground zero, and smaller amounts (30 mR h1) over Ely,
to a HYSPLIT compatible form and are available to other HYSPLIT Nevada, 241 km northeast of ground zero (Quinn et al., 1986).
users as monthly les (http://www.ready.noaa.gov/archives.php). This was the rst case of observed signicant radiation over
The data are on a 2.5 global latitudeelongitude grid with a tem- populated areas since the Trinity test in New Mexico in July
poral resolution of 6 h and are on 17 pressure levels between 1945.
1000 hPa and 10 hPa. Given that the spatial and temporal resolu- Winds at the surface were calm at detonation time, 18 m s1
tions of the NNRP data are rather coarse, HYSPLIT model results will from the south at 3000 m MSL, and 34 m s1 from the south-
also be compared with simulations using higher resolution mete- southwest at 5800 m MSL. There was very little directional shear
orology (12 km) from the Weather Research and Forecasting (WRF) throughout the column, however there was a strong speed
model in Section 5 to mimic the resolution of the regional forecast discontinuity at approximately 1700 m MSL (Hawthorne, 1979)
models used in operations today. with calm winds below and winds greater than 10 m s1 from the
south-southwest above this height. The surface winds increased
3. Nevada nuclear tests used in study throughout the day with reports of gusts to 25 m s1 at Ely, NV. In
addition, there were several reports of light precipitation over parts
The United States conducted a total of 928 atmospheric and of northern Nevada and northern Utah at some point on March 7
underground nuclear tests at the Nevada Test Site between 1951 and March 8 associated with a stationary front and a low pressure
and 1992 (Fehner and Gosling, 2006); 120 mostly atmospheric tests system in the area.
were conducted between 1951 and 1958. The Nevada Test Site is
owned and operated by the Department of Energy and is located in 3.3. Upshot-Knothole Annie shot
the southern part of the Great Basin, approximately 90 miles
northwest of Las Vegas, Nevada. The site covers approximately 1375 The third weapons test selected for this study was the rst of
square miles of dessert with elevations ranging between 3280 and eleven shots in the Upshot-Knothole Operation. The Annie shot was
7675 feet above mean sea level (MSL). detonated from a 91 m tower at Yucca Flat on March 17, 1953, at
The choice of which nuclear tests to simulate was based on 1320 UTC and had a yield of 16 kT. The cloud top was observed near
those that were detonated close to (but above) the surface and, for 12,500 m MSL. This test, broadcast live over radio and television,
all but one, those that had signicant doses to far downwind lo- was designed to determine the effects of a nuclear blast on houses,
cations. The exception was for the Buster-Jangle Sugar shot, which automobiles, mannequins, and other man-made materials for civil
46 G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55

defense purposes and was also open to select members of the press, eastern portion of the observed pattern may have been deposited in
the public, and local government ofcials. rainfall.
The observed wind directions at the time of detonation were Harry produced signicant fallout to the east over St. George,
consistently from 240 to 290 between the surface and 15,000 m Utah, when the winds shifted slightly north of where they were
MSL with a strong gradient of wind speed between 2500 m and predicted, prompting ofcials to recommend sheltering by the
2700 m MSL (Hawthorne, 1979). A cold front was moving south public during the cloud passage and increased the publics concern
toward the Nevada Test Site on the morning of the test from central of future above ground nuclear testing. A maximum exposure rate
Nevada and Utah (Steadman et al., 1983); however no precipitation of 350 mR h1 was observed between 3 and 4 h after detonation in
was reported in the area with this frontal system. The result was a St. George. Unfortunately, the radioactivity contours could not be
rather narrow plume that was observed to the east of the test site closed to the north and south of the main axis of heavy deposition
across southern Nevada and Utah. St. George, Utah reported a due to limited observations. This reduced the area available in this
maximum exposure rate of 26 mR h1 at approximately 3 h after study for direct comparison between the observed and modeled
time zero. From the fallout pattern it was determined that the plumes.
lower portions of the debris cloud took a more southeasterly di-
rection than the upper portions of the cloud causing two axiss of 3.6. Plumbbob Smoky shot
maximum dose rate (Steadman et al., 1983).
The Smoky shot, the fteenth from the Pumbbob Operation, had
a yield of 44 kT and was detonated from a 195 m tower at 1230 UTC
3.4. Upshot-Knothole Simon shot
on August 31, 1957. This tower was higher than previous shots in
attempt to lower the fallout immediately downwind of ground
The Simon shot from the same Upshot-Knothole Operation as
zero. In addition, this test was postponed several times in order to
the Annie shot, occurred on April 25, 1953, at 1230 UTC, and was the
get the best weather conditions (Harris et al., 1981). The cloud
largest test at that time and one of the largest near-surface deto-
extended to a height of approximately 12,000 m MSL.
nations ever at the Nevada Test Site with a yield of 43 kT; much
Radioactive fallout was very heavy southeast of ground zero
higher than was anticipated. Like the Annie shot, the Simon
with peak exposure rates up to 400 mR h1 (Quinn et al., 1982),
weapon was detonated from a 91 m tower and the cloud extended
which was only slightly higher than expected although the direc-
to a height of approximately 13,000 m MSL. Unfortunately, many
tion of the plume axis was further southeast than predicted. The
military personnel involved with this test in trenches near ground
radioactive deposition pattern extended initially east-southeast to
zero were exposed to radiation levels of up to 100 R h1 immedi-
the Arizona/Utah border and then turned to the northeast into
ately after the blast (Ponton et al., 1982b). Heavy fallout was re-
Utah. Surface winds at the time of detonation were calm below
ported east of ground zero in a narrow corridor across southern
1580 m MSL (burst height); from the north up to 4500 m MSL with
Nevada and northwestern Arizona with dose rate readings of
speeds between 2 and 5 m s1, and from the west-northwest above
0.46 R h1 reported on highways in the region leading to several
4500 m MSL with speeds between 6 and 20 m s1. No precipitation
road closures (Steadman, 1988; Fehner and Gosling, 2006). The day
was observed at the time of detonation near ground zero, however
after the detonation, Troy, New York, reported an integrated dose of
precipitation was observed 350 km east at Bryce Canyon and
2 R attributed to the Simon shot from a heavy rain shower.
Marysvale, in south central Utah, on August 31 and September 1
Surface winds were from the north-northwest at 3 m s1 at the
and may have contributed to the deposition in Utah and further
time of detonation; however, a column of light winds from the
northeast in Wyoming (note that no precipitation was observed in
northeast was observed just above the surface to approximately
the NNRP data).
2500 m MSL (Hawthorne, 1979). Above 2500 m MSL, the winds
were constant from the west-northwest, increasing in speed with
4. Model conguration and results using global NNRP data
height. No precipitation was observed in the area.
It should also be noted that the contour patterns were con-
The HYSPLIT model was congured to simulate the transport,
structed with a large degree of uncertainty (Steadman, 1988),
dispersion, and deposition of nuclear materials from the detona-
especially in northwestern Arizona, due to very limited and widely
tions of six nuclear tests with yields ranging from 1.2 to 44 kT,
separated observations of radioactivity, so comparisons with model
which implied stabilized cloud tops of between 3700 and 12,500 m
results will be subject to this uncertainty as well. The 100 mR h1
above ground level from Table 1. The model was run using the
contour, however, is believed to be fairly accurate based upon
global NNRP data for each of ve particle distributions (Fig. 2) for a
ground and aerial data in the area.
24-hour duration to produce concentration and deposition on two
primary output grids (Table 3) centered over the test site; a 0.02
3.5. Upshot-Knothole Harry shot horizontal resolution ne grid covering a 10  10 area, and a 0.08
horizontal resolution coarse grid covering a 30  60 area. Given
The ninth shot of the Upshot-Knothole series, Harry, produced a the longer range transport of the higher yield weapons, the coarser
yield of 32 kT from a detonation on a 91 m tower on May 19, 1953, resolution grid will be used in the remainder of the discussion of
at 1205 UTC (Hawthorne, 1979). The debris cloud height reached model results. In addition, the model integration time step was set
approximately 13,000 m MSL. Surface winds at the time of deto- to one minute to ensure consistency of the comparisons between
nation were from the north-northeast at 7 m s1. However, from the simulations.
the height of the burst up to 4500 m MSL, winds were from the After each HYSPLIT simulation completed, the resulting
southwest at speeds between 7 and 13 m s1. Above 4500 m MSL, deposition was converted to dose rate with a xed decay time of
winds were consistently from the west-northwest at speeds be- 12 h following detonation to be able to compare to measured
tween 15 and 40 m s1. Although precipitation was reported in the dose rates. The measured data were obtained from the NOAA
downwind sector on May 19 and 20, it has been determined (Quinn Weather Service Nuclear Support Ofces (WSNSO) effort to
et al., 1981) that precipitation should not have inuenced the reevaluate the measurement data in the 1980s (Steadman et al.,
deposition of radioactive material immediately downwind of 1983; Steadman, 1988; Quinn et al., 1981, 1982) for all the deto-
ground zero, but radioactivity observed in the far northern and nations except the Sugar shot which was obtained from Nagler
G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55 47

and Telegadas (1956). The contoured plumes from each of the six pattern and in four tests the modeled pattern is either north (Annie,
shots were digitized and converted to standard vector format so Harry, Smoky) or east (Sugar) of the observed pattern. The cloud
that statistical results (Figure of Merit in Space, FMS) could be width indicated by the deposition pattern is just a reection of the
calculated based on the percent of overlap between the observed variation of wind direction with height, suggesting that the reso-
and modeled plumes. Unfortunately, some of the observed lution of the NNRP data captured this essential feature of the
plumes are very complex and do not always intersect the source meteorology.
location or extend in all directions downwind due to the un- The contamination from the Sugar test (Fig. 3a), being the rst
availability of measurements. In order to be able to provide a surface-based nuclear test in the United States, was well docu-
quantitative estimate of the overlap, only the portion of the mented within a few kilometers of ground zero. However, the
observed plume (dose contours) that had two dened edges was contours drawn farther downwind are less reliable due to a limited
considered in this evaluation. number of radiological measurements and therefore the wide
Fig. 3aef show the 12-hour modeled dose rate contours deposition pattern may be uncertain. The areas of highest dose
(mR h1) from deposition for the smallest contour used in this rates just to the north of ground zero are similar to the modeled
study using the default Glasstone particle distribution and the results and also indicate a slightly northeast orientation of the
NNRP data (dashed red contour) for each of the nuclear shots centerline, although not as much as the modeled pattern, and
overlaid with the observed dose rate contours (shaded). For refer- certainly within the limits of the resolution of any wind direction
ence, the WRF simulations are also indicated by solid black con- measurements that would have been used by the NNRP.
tours and will be discussed later. Although not shown, plots for The model was able to capture most of the area of observed
each of the other particle distributions were similar. Notice that the contamination from the Easy simulation (Fig. 3b), although it did
area covered is a fraction of the original observed plume area due to not extend as far to the north as observed. Given the relatively
discontinued contours in the observed data near the source and coarse meteorological data, the results were quite good.
downwind. To view the complete observed dose rate patterns, the Unfortunately the dose rate patterns did not extend back to
reader is referred to the documents mentioned previously. Given ground zero in the Annie (Fig. 3c) and Harry (Fig. 3e) shots, how-
the coarse resolution of the meteorological data, the results are ever it is clear that the modeled depositions in both cases had a
qualitatively very good with all of the cases showing transport in more northerly component than observed. Dose rate plots for the
the general direction of the observed deposition. Differences Annie shot produced in 1956 (Nagler and Telegadas, 1956) show a
observed occur primarily in the width and displacement of the well-dened northern extent of the plume with the centerline in an
deposition patterns; in all but one nuclear test (Sugar) the width of east-southeast direction before spreading northeast near the
the modeled deposition pattern is comparable to the observed Nevada/Utah border. The transport to the southeast could be due to

Fig. 3. HYSPLIT calculated plume using NNRP (dashed red contour) and WRF meteorological data (solid black contour) overlaid onto the observed plume (gray shaded) for the A)
Sugar (0.4 mR h1 contour), B) Easy (4 mR h1 contour), C) Annie (0.1 mR h1 contour), D) Simon (4 mR h1 contour), E) Harry (4 mR h1 contour), and F) Smoky (1 mR h1 contour)
shots using the Glasstone particle distribution. Ground zero is indicated by GZ. (For interpretation of the references to colour in this gure legend, the reader is referred to the web
version of this article.)
48 G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55

the lighter observed winds in the lower layers from southwest to simulations all had trouble getting the initial southeasterly move-
northwest that the NNRP data did not capture. The modeled ment of the plume.
pattern for Harry may have been shifted slightly north of the Fig. 4aef present Figure of Merit in Space (FMS, Warner et al.,
observed pattern due to less ow from the west-northwest than 2003) scores for each shot based on the observed contour value
was observed at mid-levels above 5000 m. and the particle distribution type. The FMS score provides a
Finally, the modeled deposition pattern of the Smoky simulation quantitative measure of the overlap of the modeled and observed
(Fig. 3f), although very similar in size and shape, completely missed plume segments. The denition of FMS, expressed as a percentage,
the area of observed deposition by passing to the north. The is the ratio of the intersection of the modeled (A1) and the observed
observed pattern had a very sharp northern extent with the bulk of (A2) areas to the union of the areas.
the deposition extending east-southeast before curving to the
northeast near the Nevada/Utah border. As mentioned earlier, the A XA2
Figure of Merit in Space FMS 100$ 1 (1)
observed plume was actually farther southeast than predicted at A1 WA2
the time of detonation. The meteorological situation was rather
complex with a low pressure area over southwestern Utah and an A value near 100 is an indication of a good match between the
upper-level trough passing through the region, which contributed two patterns and hence good model performance, however a small
to the strong ow from the west-northwest throughout the col- number does not necessarily mean low model performance as the
umn. The NNRP data had mostly westerly ow at mid-levels and model plume may be identical in size and shape to the observed
light and variable winds below 3000 m. Similar results were plume but shifted a few degrees and therefore giving a low FMS
recently found by Schoeld (2012) in that HYSPLIT, HPAC and the score. Thus the need to show the plume shapes in addition to
Weather Research & Forecasting with Chemistry (WRF-CHEM) presenting the FMS scores.

Fig. 4. Figure of Merit in Space (FMS) scores (%) for the Sugar (A), Easy (B), Annie (C), Simon (D), Harry (E), and Smoky (F) nuclear tests using the NNRP meteorological data.
G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55 49

In general, FMS scores decrease with increasing dose rate con- further down-wind, however the horizontal spread of the plume
tours with the exception of the Sugar (Fig. 4a) and Smoky (Fig. 4f) was similar to that observed. It should also be noted that because
shots which had very little overlap overall. The decreasing score the observed dose rates were obtained from contours and not point
with increasing dose rate is to be expected given the smaller measurements, the actual peak observed values are not known at
footprint of the highest dose rate contours. In regard to the particle each cross-section.
distribution type, the FMS scores are very similar for all distribution To provide a quantitative analysis of the model cross-section
types, especially at the lower dose rate contours. However, the results regardless of any mismatches in plume position with the
Glasstone distribution tends to have the highest FMS score per dose observations, the average dose rate across the plume cross-section
rate contour for all but the Smoky shot (Fig. 4f), which produced was calculated for each shot and plotted against the average
very low FMS scores at all contour levels and all distribution types observed dose rate at the same downwind distances. Fig. 6aef
due to the modeled plumes being north of the observed plumes show the average observed (hashed column) and modeled dose
with little overlap. The Glasstone particle/activity distribution is rates (mR h1) for each shot and each particle distribution along the
quite different from the others in that it has the least activity in the cross-sections using a log scale on the ordinate axis. The cross-
smallest and largest particle sizes. The highest FMS scores were sections are ordered from closest (left) to farthest (right) from
obtained from the Simon shot (Fig. 4d) with FMS scores above 70% ground zero. In general, the modeled dose rates are lower than the
at the lowest dose rate contour; a result of the very good overlap observed dose rates, especially for cross-sections farther from
between modeled and observed dose rate contours. ground zero. The Glasstone distribution tends to have the best
To determine how well the model was able to capture dose rates overall match to the observed dose rate and a larger average dose
across the plume footprint, three or four cross-sections along the rate than the other distributions, but most of the other particle
plume footprint were extracted from the model calculated dose distributions provide very similar results to each other. It can also
rates and compared with the dose rates at the observed contour be seen that the model tends to overestimate the dose rates near
intersections. For example, Fig. 5a shows the cross-sections (solid the source location for three of the six tests; however since the
vertical lines) on a map of the observed dose rates for the Simon observed peak dose rate is not known this conclusion may be
shot using the Glasstone particle distribution along with the incorrect, and in fact, the observed dose rate may actually be higher
modeled and observed dose rates across each cross-section than shown. It should also be noted that if fractionation had been
(Fig. 5bed). In this case the model under-estimated the peak dose included in the calculation of dose rates, there may have been
rate near the test site and may have over-estimated the peak greater agreement for some of the simulations as the calculated

Fig. 5. Map of observed dose rates (A) from the Simon shot with cross-section locations indicated as solid vertical lines. Modeled and observed dose rates from each cross-section
with increasing distance from ground zero: 115.83 W (B), 115.00 W (C), and 114.00 W (D).
50 G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55

Fig. 6. Average dose rates (mR h1) along cross-sections of the plume footprint for each nuclear test and particle distribution using the NNRP meteorological data: Sugar (A), Easy
(B), Annie (C), Simon (D), Harry (E), and Smoky (F).

dose rates close-in might have been slightly reduced and the dose The Advanced Research WRF (WRF-ARW; Skamarock et al.,
rates farther downwind may have been slightly increased. The large 2008), version 3.4.1, was congured with two nested domains
difference seen between the measured and modeled average dose (Fig. 7); the outermost grid at 36-km resolution covering the
rates in the Sugar shot (Fig. 6a) at the LAT38_00 cross-section is contiguous U.S., and the innermost grid at 12 km covering the
primarily due to a secondary peak in the observed dose rate in this Western and Central U.S. The initial and boundary conditions for
vicinity that the model did not capture. WRF originated from the NCEP/NCAR Global Reanalysis (NNRP;
used in the previous section), which is available every 6 h at 2.5
5. Model conguration and results using WRF-ARW data spatial resolution. Twenty-seven layers were dened for the ver-
tical grid of the WRF simulation with the rst layer top at about
In the previous section the HYSPLIT model was congured for a 32 m. The WRF physics options included the WSM 3-class simple
nuclear detonation and compared with observed dose rates from ice scheme (Dudhia, 1989) for microphysics, the Rapid Radiative
six Nevada nuclear tests using the global reanalysis meteorological Transfer Model (Mlawer et al., 1997) for longwave radiation, the
data. The model was able to reproduce similar shapes and magni- Dudhia scheme (Dudhia, 1989) for shortwave radiation, the Unied
tudes of the observed dose rate patterns even though the meteo- Noah land-surface model (Chen and Dudhia, 2001), the Yonsei
rological data was very coarse (2.5 latitude by 2.5 longitude grid University scheme (Hong et al., 2006) for PBL parameterization and
or approximately 275 km  200 km) in comparison to current the Grell-Devenyi ensemble scheme (Grell and Devenyi, 2002) for
operational meteorological forecast models (w12 km). Given that the cumulus option.
ner resolution meteorological data would be used in an IND event The WRF simulations were run with two-way nesting using the
that occurred today, the six test cases were rerun using 12 km NNRP data as initial and boundary conditions. Observational
horizontal resolution data from the Weather Research and Fore- nudging was also incorporated using upper-air soundings launched
casting (WRF) model. near ground zero near the time of detonation (Hawthorne, 1979) to
G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55 51

Fig. 7. WRF-ARW outer 36 km domain and nested 12 km domain (d02). The black dot is the location of the Nevada Test Site.

adjust predicted values toward observations at hours when data section and the results (not shown) looked very similar. However,
were available, however only the U- and V-wind components were to better discern any differences, the Fraction Bias (FB) was calcu-
nudged since neither temperature nor water mixing ratio mea- lated for each simulation using the following equation
surements were available. The Annie, Easy, Harry, and Simon tests  
used one sounding near the time of detonation whereas the Smoky Cp  Co
Fractional Bias FB 2  (2)
test had two soundings three hours apart and the Easy test had Cp Co
three soundings in the rst four hours after detonation.
As discussed in the previous section, Fig. 3aef show the HYSPLIT where Co and Cp are the observed and predicted average dose rates,
calculated plume using the NNRP (dashed red contour) and the respectively. Fig. 9(aee) shows scatter plots of the absolute value of
WRF (solid black contour) meteorological data for the smallest NNRP and WRF fractional biases for each particle distribution. Each
contour used in this study overlaid with the observed pattern (gray symbol represents the FB from a different cross-section of a
shading). The WRF simulations were very similar to the NNRP particular nuclear test. The Glasstone distribution has less bias than
simulations except for the Simon (Fig. 3d) test where the WRF the other particle distributions with all but four points having a FB
simulation produced a much broader plume farther downwind of less than one, whereas the other distributions have between
than was observed. WRF may have indicated more mixing than was eight and thirteen FBs above one. However, there does not appear
observed in the layer containing the nuclear debris. In all but the to be any appreciable improvement in the FB when using WRF
Simon test, using the WRF model improved the FMS scores for the versus NNRP as most values are on either side of the 1:1 line. A
lowest contour level by 2e7%. review of the FB values (not shown) shows that in four of six tests a
The HYSPLIT simulations and statistical analysis discussed in the lower FB was observed using WRF at the closest cross-section to the
last section using NRRP data (Fig. 4) were repeated using the WRF source location and higher FBs farther from the source. Improve-
meteorological data. Fig. 8aef present the FMS scores for each shot ments close to the source are expected with the WRF meteoro-
based on the contour value and the particle distribution type for the logical data since the plume should start out in a direction more
HYSPLIT calculations using WRF meteorological data. In ve of the representative of the actual ow due to the use of an observed
six simulations (Sugar, Easy, Annie, Harry, Smoky) the FMS scores sounding near the source location in the WRF nudging algorithm;
were higher (w5e20%) using the WRF data instead of NRRP data, however it is unclear why the results are lower as distance in-
especially at the higher dose rate contours. The largest increase in creases. It is possible that the vertical mixing is enhanced using the
FMS scores was observed for the Easy shot as the center of the ner vertical and horizontal resolution WRF data, which may have
plume was moved slightly north of the NRRP position and provided diluted the plume at farther distances.
a better t to the peak dose rates. The Simon shot had FMS scores of Using a ner resolution meteorological model had a positive
10e30% lower at the two lowest dose rate contours using the WRF impact in the FMS scores for most of the simulations and the
data due to a wider deposition pattern especially farther downwind average dose rates near the source locations, however improve-
and also due to a slight shift to the north of the observed deposition ments were not observed for every test and at all distances
pattern. downwind from ground zero. It is possible that if additional
The average dose rates were calculated in the same manner for soundings were available across the modeling domain before and
the WRF calculations as was done for NNRP (Fig. 6) at each cross- after detonation that results may have improved even more,
52 G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55

Fig. 8. Same as Fig. 4, but using the WRF meteorological data.

especially at locations farther from the source. However, it is correlated with the measured dose rates due to the plumes being
encouraging that the model was able to reproduce the general di- shifted slightly in space, the model was within a factor of three of
rection and plume dimensions correctly even with the very coarse the measured 95th percentile dose rate as shown by the non-paired
meteorology of the NNRP, possibly indicating that the primary statistics in Table 4, with four of six cases within a factor of one. The
transport mechanism was the synoptic scale ow patterns at higher simulations using the WRF model performed better than the NNRP
levels. Of course given a different source location with complex simulations in all but the Simon test. Normalized mean square er-
terrain features or a more complicated synoptic situation may rors were also lower using the WRF data in all but the Simon and
produce better results using the ner resolution meteorological Harry tests. Finally, the fractional biases were generally less than
data. 0.3 with the exception of the Sugar and Harry tests where the
model had an overall under-prediction with fractional biases
6. Modeled dose rates for locations in the Town Data Base around 1. The Annie test was the only other simulation that the
model had an under-prediction.
As a nal measure of model performance, dose rates and plume Fig. 10 shows the scatter plots of the measured versus modeled
arrival times were calculated for those locations in the Town Data time of arrival for the six tests for all locations that had a non-zero
Base (Thompson et al., 1994) having non-zero measured dose rates value for modeled dose rate (lled circles represent the NNRP
for each of the six nuclear tests. The Town Data Base, developed as simulations and triangles represent the WRF simulations). In all but
part of the U.S. Department of Energys Off-Site Radiation Exposure the Annie and Smoky simulations, the results show a high corre-
Review Project (ORERP), contains estimated fallout arrival times lation between the modeled and measured times of plume arrival,
and H12 exposure rates for select locations in Arizona, California, and except for a few locations with a long modeled time of arrival
Nevada, and Utah for 74 nuclear events, including the six tests used (20 h), the Smoky test results would also be well correlated. The
in this study. Although the calculated dose rates were not well model tended to have slower transport speed for the Annie test,
G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55 53

Fig. 9. Scatter plots of the absolute value of Fractional Bias (FB) between simulations run using NNRP and WRF meteorological data for each particle distribution simulated by
HYSPLIT (A-Glasstone, B-Izrael, C-Small boy, D-Heffter, E-DELFIC).

especially when the WRF meteorological data was used in the 7. Summary
simulation indicating that the actual fallout may have come from a
different height than the modeled height or the winds were under- The NOAA Air Resources Laboratorys (ARL) Hybrid Single Par-
estimated by the model. The WRF simulations improved the results ticle Lagrangian Integrated Trajectory (HYSPLIT) model, the pri-
for all simulations except for the Annie and Smoky tests. The time of mary atmospheric and dispersion model in use today by NOAA,
arrival calculations for the Harry test also compare favorably with was congured to simulate the dispersion and deposition of nu-
results obtained by Moroz et al. (2010) and Cederwall and Peterson clear materials from a surface-based nuclear detonation using
(1990) where they compared modeled time of arrival and exposure publicly available information on nuclear explosions. The model
rates with those obtained from gummed-lm deposition mea- was evaluated against the measurements of nuclear fallout from
surements (Beck, 1984, 1990). However, as Moroz et al. (2010) six nuclear tests conducted between 1951 and 1957 at the Nevada
found, the fallout pattern (not shown) was not as broad as found Test Site using the global NCEP/NCAR Reanalysis Project (NNRP)
by Cederwall and Peterson (1990), and the modeled area of highest and the Weather Research and Forecasting (WRF) meteorological
deposition was shifted slightly north of St. George, UT, missing the data as input. Although four other distributions were considered in
high dose rate measured there. the analysis, the information on the nuclear cloud characteristics
and the default particle sizes and activity distributions from The
Effects of Nuclear Weapons by Glasstone and Dolan (1977) tended
Table 4 to produce the best results for all tests and all dose rate contours.
Statistical results from a comparison of measured versus calculated dose rates for The model was able to reproduce the general direction and
locations in the Town Data Base (Thompson et al., 1994), including the 95th deposition patterns using the coarse NNRP data with Figure of
percentile values, the normalized mean square error (NMSE) and the fractional bias Merit in Space (FMS) scores in excess of 50% for four of six sim-
(FB).
ulations for the smallest dose rate contours, with FMS scores
Sugar Easy Annie Simon Harry Smoky declining for higher dose rate contours. When WRF meteorological
No. of locations 40 34 40 55 85 95 data were used the FMS scores were 5e20% higher in ve of the
95th % Measured 4.73 9.97 25.26 31.00 74.37 25.56 six simulations, especially at the higher dose rate contours. When
95th % NNRP 2.13 13.55 39.70 *67.12 28.20 42.01 compared with measurements of dose rate and time of arrival
95th % WRF *3.19 *12.16 *25.26 99.61 *30.27 *23.82
from the Town Data Base, similar results were found with the WRF
NMSE NNRP 6.6 2.9 7.7 *10.2 4.4 19.9
NMSE WRF *3.8 *2.8 *5.3 17.3 4.4 *17.2 simulations providing better results for four of six simulations.
FB NNRP 1.04 *0.04 0.19 *0.27 *e0.99 0.31 Given the result that the model was able to reproduce the general
FB WRF *0.89 0.13 0.19 0.71 1.05 *0.04 direction and magnitude of the observed dose rate patterns of
A * indicates the model run with the best performance when using either the these six nuclear tests even with relatively coarse meteorological
NNRP or WRF data. data, this simulation will be added to the tools available to
54 G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55

Fig. 10. Scatter plots of modeled versus measured time of arrival (h) for the six tests for all locations that had a non-zero value for modeled dose rate. The lled circles represent the
NNRP simulations and the triangles represent the WRF simulations.

researchers and forecasters for simulating the transport and Draxler, R.R., 1982. Measuring and modeling the transport and dispersion of
kRYPTON-85 1500 km from a point source. Atmos. Env. (1967) 16 (12), 2763e
deposition from small nuclear devices using the HYSPLIT model.
2776. http://dx.doi.org/10.1016/0004-6981(82)90027-0.
With even higher resolution meteorological data now being Draxler, R.R., Hess, G.D., 1997. Description of the HYSPLIT_4 Modeling System. NOAA
implemented into U.S. National Weather Service operations (4 km Tech. Memo. ERL-ARL-224. NOAA/Air Resources Laboratory, Silver Spring, MD.
CONUS NEST; Ferrier et al., 2011; Janjic and Gall, 2012), with Available online at: http://www.arl.noaa.gov/TechMemos.php.
Draxler, R.R., Hess, G.D., 1998. An overview of the HYSPLIT_4 modeling system of
improved resolution of local-scale ow patterns, it is hoped that trajectories, dispersion, and deposition. Aust. Meteor. Mag. 47, 295e308.
further improvements will be observed in HYSPLIT model results Draxler, R.R., 1999. HYSPLIT4 Users Guide. NOAA Tech. Memo. ERL ARL-230. NOAA/
in the near future. Air Resources Laboratory, Silver Spring, MD. Available online at: http://www.arl.
noaa.gov/TechMemos.php.
Draxler, R.R., Ginoux, P., Stein, A.F., 2010. An empirically derived emission algorithm
Acknowledgments for wind blown dust. J. Geophys. Res. Atmos. 115, D16212. http://dx.doi.org/10.
1029/2009JD013167.
Draxler, R.R., Rolph, G.D., 2012. Evaluation of the Transfer Coefcient Matrix (TCM)
The authors would like to thank Jerome L. Heffter and Dr. Vin- approach to model the atmospheric radionuclide air concentrations from
cent J. Jodoin for their constructive comments and suggestions Fukushima. J. Geophys. Res. Atmos 117. http://dx.doi.org/10.1029/2011JD017205.
during the development of the model. DTRA, 2001. The HPAC Users Guide, Version 4.0.3. Prepared for Defense Threat
Reduction Agency, Contract DSWA01-98-C-0110, by Science Applications In-
ternational Corporation, Rep. HPAC-UGUIDE-02-U-RAC0.
References Dudhia, J., 1989. Numerical study of convection observed during the winter monsoon
experiment using a mesoscale two-dimensional model. J. Atmos. Sci. 46, 3077e
3107. http://dx.doi.org/10.1175/1520-0469 (1989)046<3077:NSOCOD>2.0.CO;2.
Bartnicki, J., Saltbones, J., 2008. Modelling atmospheric dispersion of radioactive
Eckerman, K.F., Leggett, R.W., 1996. DCFPAK: Dose Coefcient Data File Package for
debris released in case of a nuclear explosion using the Norwegian SNAP model.
Sandia National Laboratory. ORNL/TM-13347. Oak Ridge National Laboratory,
Croat. Meteorol. J. 43, 111e115. Available online at: http://www.meteohmd.hr/
Oak Ridge, TN.
izdanja/hmc/43/WEB/Oral.htm.
Beck, H.L., 1984. Estimates of Fallout From Nevada Weapons Testing in the Western England, T.R., Rider, B.F., 1993. Evaluation and Compilation of Fission Yields. LA-UR-
94e3106; ENDF-349. Los Alamos National Laboratory.
United States Based on Gummed-lm Monitoring Data. No. EML-433. USDOE
Environmental Measurements Lab., New York. Fehner, T.R., Gosling, F.G., 2006. Atmospheric Nuclear Weapons Testing, 1951e1963.
Battleeld of the Cold War: The Nevada Test Site, Volume 1 (pdf). DOE/MA-
Beck, H.L., Helfer, I.K., Bouville, A., Dreicer, M., 1990. Estimates of fallout in the
0003. Department of Energy, Washington, D.C.
continental US from Nevada weapons testing based on gummed-lm moni-
Ferber, G.J., Heffter, J.L., 1961. A Comparison of Fallout Model Predictions with a
toring data. Health Phys. 59 (5), 565e576.
Consideration of Wind Effects et seq. AEC TID-7632, 122.
Cederwall, R.T., Peterson, K.R., 1990. Meteorological modeling of arrival and depo-
sition of fallout at intermediate distances downwind of the Nevada Test Site. Ferber, G.J., Telegadas, K., Heffter, J.L., Smith, M.E., 1976. Air concentrations of
Krypton-85 in the midwest United States during January-May 1974. Atmos.
Health Phys. 59 (5), 593e601.
Environ. 11, 379e385. http://dx.doi.org/10.1016/0004-6981(77)90168-8.
Chang, C.C., Hanna, S.R., Boybeyi, Z., Franzese, P., 2005. Use of Salt Lake City URBAN
Ferrier, B.S., Wang, W., Colon, E., 2011. Evaluating cloud microphysics schemes in
2000 eld data to evaluate the Urban Hazard Prediction Assessment Capability
nested NMMB forecasts. In: Preprints, 24th Conf. on Wea. Fore./20th Conf. on
(HPAC) dispersion model. J. Appl. Meteor. 44, 485e501. http://dx.doi.org/10.
1175/JAM2205.1. Num. Wea. Pred. Amer. Meteor. Soc., Seattle, WA, 14B.1. Available online at:
https://ams.confex.com/ams/91Annual/webprogram/Paper179488.html.
Chen, F., Dudhia, J., 2001. Coupling and advanced land surface-hydrology model
with the Penn State-NCAR MM5 modeling system. Part I: model implementa- Glasstone, S., Dolan, P.J., 1977. The Effects of Nuclear Weapons. Published by the U.S,
third ed. Department of Defense and the U.S. Department of Energy. http://dx.
tion and sensitivity. Mon. Wea. Rev. 129, 569e585. http://dx.doi.org/10.1175/
doi.org/10.2172/972902.
1520-0493 (2001)129<0569:CAALSH>2.0.CO;2.
G.D. Rolph et al. / Journal of Environmental Radioactivity 136 (2014) 41e55 55

Grell, G.A., Devenyi, D., 2002. A generalized approach to parameterizing convection NATO, 2010. Warning and Reporting and Hazard Prediction of Chemical, Biological,
combining ensemble and data assimilation techniques. Geophys Res. Lett. 29, Radiological and Nuclear Incidents (Operators Manual). NATO/PfP UNCLASSI-
1693. http://dx.doi.org/10.1029/2002GL015311. FIED, ATP e 45 (D). http://www.assistdocs.com/search/document_details.cfm?
Harris, P.S., Lowery, C., Nelson, A.G., Obermiller, S., Ozeroff, W.J., Weary, E., 1981. Shot ident_number97673.
SMOKY, A Test of the PLUMBBOB Series (DNA 6004F). Defense Nuclear Agency. Norment, H.G., 1979a. DELFIC: Department of Defense Fallout Prediction System,
Continental U.S. Tests (DNA 1251-1-FX). In: Hawthorne, H.A. (Ed.), 1979. Compila- vol. I. Fundamentals. ADA 088e367.
tion of Local Fallout Data from Test Detonations 1945e1962 Extracted from Norment, H.G., 1979b. DELFIC: Department of Defense Fallout Prediction System,
DASA 1251, vol. I. Defense Nuclear Agency. vol. II. Users Manual. ADA 088e512.
Heffter, J.L., 1969. ARL Fallout Prediction Technique. ESSA Tech. Memo. ERLTM-ARL Ponton, J., Rohrer, S., Maag, C., Massie, J., 1982a. Shots SUGAR and UNCLE, the Final
13. NOAA/Air Resources Laboratory, Silver Spring, MD. Available online at: Tests of the BUSTER-JANGLE Series (DNA 6025F). Defense Nuclear Agency.
http://www.arl.noaa.gov/TechMemos.php. Ponton, J., Shepanek, R., Massie, J., Rohrer, S., Maag, C., 1982b. Operation Upshot-
Heffter, J.L., 1980. Air Resources Laboratories Atmospheric Transport and Dispersion Knothole, 1953 (DNA 6014F). JRB Associates, Inc., McLean, VA.
Model (ARL-ATAD). NOAA Tech. Memo. ERL ARL 81. NOAA/Air Resources Lab- Quinn, V.E., Urban, V.D., Kennedy, N.C., 1981. Analysis of Operation Upshot-Knothole
oratory, Silver Spring, MD. Available online at: http://www.arl.noaa.gov/ 9 (HARRY) Radiological and Meteorological Data. NVO-233. National Oceanic
TechMemos.php. and Atmospheric Administration, Weather Service Nuclear Support Ofce, Las
Hicks, H.G., 1981. Results of Calculations of External Radiation Exposure Rates from Vegas, Nevada.
Fallout and the Related Radionuclide Composition. UCRL-53152, parts 1e8, Quinn, V.E., Kennedy, N.C., Urban, V.D., 1982. Analysis of Operation Plumbbob Nu-
1981. Lawrence Livermore National Laboratory, Livermore, CA. clear Test SMOKY Radiological and Meteorological Data. NVO-249. National
Hicks, H.G., 1984. Results of Calculations of External Gamma Radiation Exposure Oceanic and Atmospheric Administration, Weather Service Nuclear Support
Rates From Local Fallout and the Related Radionuclide Compositions of Selected Ofce, Las Vegas, Nevada.
U.S. Pacic Events. Lawrence Livermore National Laboratory, Livermore, CA. Quinn, V.E., Urban, V.D., Kennedy, N.C., 1986. Analysis of Operation Tumbler-
UCRL-53505. Snapper Nuclear Test EASY Radiological and Meteorological Data. NVO-297.
Hoecker, W.H., Machta, L., 1990. Meteorological modeling of radioiodine transport National Oceanic and Atmospheric Administration, Weather Service Nuclear
and deposition within the continental United States. Health Phys. 59 (5), 603e Support Ofce, Las Vegas, Nevada.
617. Rolph, G.D., Draxler, R.R., Stein, A.F., Taylor, A., Ruminski, M.G., Kondragunta, S.,
Homann, S.G., Aluzzi, F., 2013. HotSpot Health Physics Codes: Version 3.0 Users Zeng, J., Huang, H., Manikan, G., McQueen, J.T., Davidson, P.M., 2009. Description
Guide. LLNL-SM-636474. National Atmospheric Release Advisory Center, Liv- and verication of the NOAA smoke forecasting system: the 2007 re season.
ermore, CA. Weather Forecast. 24, 361e378. http://dx.doi.org/10.1175/2008WAF2222165.1.
Hong, S.Y., Noh, Y., Dudhia, J., 2006. A new vertical diffusion package with an Schoeld, J.C., 2012. Mapping Nuclear Fallout Using the Weather Research & Fore-
explicit treatment of entrainment processes. Mon. Wea. Rev. 134, 2318e2341. casting (WRF) Model (No. AFIT/CWMD/ENP/12-S01). Air Force Inst. of Tech..
http://dx.doi.org/10.1175/MWR3199.1. Graduate School of Engineering and Management, Wright-Patterson AFB, OH.
Izrael, Y.A., 2002. Radioactive Fallout after Nuclear Explosions and Accidents, Skamarock, W.C., Klemp, J.B., Dudhia, J., Gill, D.O., Barker, D.M., Huang, X., Wang, W.,
Radioactivity in the Environment Series. Elsevier Science, Amsterdam, p. 3. Powers, J.G., 2008. A Description of the Advanced Research WRF Version 3.
Janjic, Z., Gall, R.L., 2012. Scientic Documentation of the NCEP Nonhydrostatic NCAR Tech Note NCAR/TN-475STR.
Multiscale Model on the B Grid (NMMB). Part 1 Dynamics. NCAR Technical Note Steadman, C.R., Kennedy, N.C., Quinn, V.E., 1983. Analysis of Upshot-Knothole 1
NCAR/TN-489STR. http://dx.doi.org/10.5065/D6WH2MZX. (ANNIE) Radiological and Meteorological Data. NVO-254. National Oceanic and
Jodoin, V.J., 1994. Nuclear Cloud Rise and Growth. AD B007 607. Air Force Institute of Atmospheric Administration, Weather Service Nuclear Support Ofce, Las
Technology, Wright-Patterson AFB, OH. Vegas, Nevada.
Kalnay, E., Kanamitsu, M., Kistler, R., Collins, W., Deaven, D., Gandin, L., Iredell, M., Steadman, C.R., 1988. Analysis of Upshot-Knothole Nuclear Test SIMON Radiological
Saha, S., White, G., Woollen, J., Zhu, Y., Chelliah, M., Ebisuzaki, W., Higgins, W., and Meteorological Data. NVO-315. National Oceanic and Atmospheric
Janowiak, J., Mo, K.C., Ropelewski, C., Wang, J., Leetmaa, A., Reynolds, R., Administration, Weather Service Nuclear Support Ofce, Las Vegas, Nevada.
Jenne, R., Joseph, D., 1996. The NCEP/NCAR 40-year reanalysis project. Bull. Am. Stein, A.F., Rolph, G.D., Draxler, R.R., Stunder, B., 2009. Verication of the NOAA
Meteor. Soc. 77, 437e470. http://dx.doi.org/10.1175/1520-0477 (1996) smoke forecasting system: model sensitivity to the injection height. Weather
077<0437:TNYRP>2.0.CO;2. Forecast. 24, 379e394. http://dx.doi.org/10.1175/2008WAF2222166.1.
List, R.J., Machta, L., Telegadas, K., 1961. Fallout from the 1961 soviet test series. Stunder, B.J.B., Heffter, J.L., Draxler, R.R., 2007. Airborne volcanic ash forecast area
Weatherwise 14 (6), 219e223. http://dx.doi.org/10.1080/00431672.1961. reliability. Weather Forecast. 22, 1132e1139. http://dx.doi.org/10.1175/
9930027. WAF1042.1.
Machta, L., List, R.J., 1956. World-wide travel of atomic debris. Science (United Telegadas, K., List, R.J., 1964. Global history of the 1958 nuclear debris and its
States) 124. meteorological implications. J. Geophys. Res. 69 (22), 4741e4753. http://dx.doi.
Machta, L., List, R.J., Telegadas, K., 1962. A survey of radioactive fallout from nuclear org/10.1029/JZ069i022p04741.
tests. J. Geophys. Res. 67 (4), 1389e1400. http://dx.doi.org/10.1029/ Telegadas, K., Ferber, G.J., Heffter, J.L., Draxler, R.R., 1978. Calculated and observed
JZ067i004p01389. seasonal and annual Kr-85 concentrations at 30e150 km from a point source.
Machta, L., Heffter, J.L., 1986. Chernobyl: the meteorological viewpoint. Environ- Atmos. Environ. 12, 1769e1775. http://dx.doi.org/10.1016/0004-6981(78)
ment (USA) 28 (7). 90325-6.
Machta, L., 1992. Finding the site of the rst Soviet nuclear test in 1949. Bull. Amer. Telegadas, K., 1979. Estimation of maximum credible atmospheric radioactivity
Meteor. Soc. 73, 1797e1806. http://dx.doi.org/10.1175/1520-0477 (1992) concentrations and dose rates from nuclear tests. Atmos. Environ. 13, 327e334.
073<1797:FTSOTF>2.0.CO;2. http://dx.doi.org/10.1016/0004-6981(79)90176-8.
Miller, C.F., Sartor, J.D., 1965. Small Boy shot fallout research program. In: Thompson, C.B., McArthur, R.D., Hutchinson, S.W., 1994. Development of the Town
Klement Jr., A.W. (Ed.), Radioactive Fallout from Nuclear Weapons Tests. U.S. Data Base: Estimates of Exposure Rates and Times of Fallout Arrival Near the
Atomic Energy Commission, Germantown, MD, pp. 44e71. Nevada Test Site. Report DOE/NV-374, U.S.. Department of Energy, Nevada
Mlawer, E.J., Taubman, S.J., Brown, P.D., Iacono, M.J., Clough, S.A., 1997. Radiative Operations Ofce, Las Vegas.
transfer for inhomogeneous atmosphere: RTTM, a validated correlated-k model Trabalka, J.R., Kocher, D.C., 2007. Bounding Analysis of Effects of Fractionation of
for the longwave. J. Geophys. Res. 102, 16663e16682. http://dx.doi.org/10.1029/ Radionuclides in Fallout on Estimation of Doses to Atomic Veterans. DTRA-TR-
97JD00237. 07e15. SENES Oak Ridge, Inc., Oak Ridge, TN, and Defense Threat Reduction
Moroz, B.E., Beck, H.L., Bouville, A., Simon, S.L., 2010. Predictions of dispersion and Agency, Fort Belvoir, VA.
deposition of fallout from nuclear testing using the NOAA-HYSPLIT meteoro- Warner, S., Platt, N., Heagy, J.F., 2003. User-oriented two-dimensional measure of
logical model. Health Phys. 99 (2), 252e269. http://dx.doi.org/10.1097/HP. effectiveness for the evaluation of transport and dispersion models. J. Appl.
0b013e3181b43697. Meteor. 43, 58e73. http://dx.doi.org/10.1175/1520-0450 (2004)043<0058:
Nagler, K.M., Telegadas, K., 1956. The Distribution of Signicant Fallout From Nevada UTMOEF>2.0.CO;2.
Tests. Prepared for the Albuquerque Operations Ofce of the U.S.. Atomic En-
ergy Commission. U.S. Weather Bureau, Washington, D.C.

S-ar putea să vă placă și