Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110, A09S40, doi:10.

1029/2004JA010985, 2005

Global thermospheric neutral density and


wind response to the severe 2003 geomagnetic
storms from CHAMP accelerometer data
E. K. Sutton, J. M. Forbes, and R. S. Nerem
Department of Aerospace Engineering Sciences, University of Colorado, Boulder, Colorado, USA

Received 18 December 2004; revised 3 May 2005; accepted 17 June 2005; published 23 September 2005.
[1] Measurements of atmospheric density near 410 km from the STAR accelerometer
on the CHAMP satellite are used to illustrate the spatial-temporal dependence of the
thermospheric response to the severe solar storms occurring during 29 October to
1 November 2003. This interval includes periods of elevated magnetic activity with KP
values of 5–9, as well as undisturbed intervals that serve to define quiet time baseline
densities. Measurements are available from 87 to +87 latitude during both day and
night at local times near 1300 and 0100 hours, respectively. During times of
maximum geomagnetic activity for this study, density measurements exhibit
enhancements of 200–300%. Northern Hemisphere daytime responses are much larger
than in the Southern Hemisphere; the origins of this effect are unknown. Nighttime
density disturbances more readily propagate to equatorial latitudes, possibly facilitated
by the predominant equatorward flow in both hemispheres due to the diurnal tides
driven by in situ EUV heating. The CHAMP density measurements are compared with
density predictions from the NRL-MSISe00 empirical density model and demonstrate
some model shortcomings. Measurements of cross-track accelerations provide the
opportunity to estimate zonal winds from the equator to about ±60 latitude,
transitioning to a measure of purely meridional winds at the turning point of the orbit
near ±87 latitude. A periodic variation in cross-track winds with an apparent period of
24 hours appears at high latitudes and exhibits similar amplitudes and temporal-
latitudinal structures to the empirical HWM-93 wind model when projected into the
cross-track direction. This periodicity is due to the displacement of geomagnetic and
geographic coordinates. At low latitudes, CHAMP and HWM-93 both yield westward
winds of order 100 ms1 during midday under quiet magnetic conditions; however,
during severely disturbed periods the HWM-93 winds generally show a greater
westward intensification (to 250 ms1) than the CHAMP measurements. At night,
CHAMP winds are near zero under quiet conditions whereas HWM-93 indicates
eastward winds of order 50–100 ms1. Under disturbed conditions the CHAMP winds
shift to westward values of order 200 to 250 ms1, while HMW-93 values do not
exceed about 50 ms1 in the westward direction. The physical origins of the observed
effects are difficult to isolate, and unequivocal interpretation will require sophisticated
numerical modeling taking into account self-consistent interactions between the
neutral winds, drifts, and ionization densities.
Citation: Sutton, E. K., J. M. Forbes, and R. S. Nerem (2005), Global thermospheric neutral density and wind response to the severe
2003 geomagnetic storms from CHAMP accelerometer data, J. Geophys. Res., 110, A09S40, doi:10.1029/2004JA010985.

1. Introduction on 29 and 30 October. At 0611 UT on 29 October, the CME


associated with an X17 flare impacted the Earth’s magnetic
[2] During the period spanning from late October to early
field. The geomagnetic aP index reached 400 for a short
November of 2003, an active region on the Sun (Active
period of time after the impact. A northward interplanetary
Region 486) produced flares and coronal mass ejections
magnetic field (IMF) Bz component was responsible for a
(CMEs) of intensities rarely seen before. Two major geo-
slight lull in geomagnetic activity from 0900 to 1800 UT on
magnetic storms were caused by CMEs arriving at the Earth
29 October. Near the end of this period, a strong southward
IMF Bz component was observed, finishing out the day with
Copyright 2005 by the American Geophysical Union. severe geomagnetic activity. The second severe storm
0148-0227/05/2004JA010985 occurred at 1600 UT on 30 October when the CME

A09S40 1 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

measurements of thermospheric total density and wind


speed derived from accelerometer measurements. Acceler-
ometer measurements have been used in the recent past to
estimate both total density [Berger and Barlier, 1981;
Forbes et al., 1996; Bruinsma et al., 2004] and wind speed
[Marcos and Forbes, 1985; Forbes et al., 1993] for quiet
and active conditions. Total density measurements represent
one measure of the global thermospheric response to
geomagnetic storms.
[5] The German satellite CHAMP (Challenging Minisa-
tellite Payload) was launched into a circular orbit on 15 July
2000. The project plan includes a 5-year mission duration to
study the gravity and magnetic fields of the Earth, with a
secondary goal of studying the upper atmosphere [Reigber
et al., 2002b]. One of the instruments on CHAMP is a
sensitive triaxial accelerometer that is capable of providing
estimates of total mass density and cross-track winds. With
a near-polar inclination, the satellite provides near-global
coverage at an approximate altitude of 410 km within two
Figure 1. Latitude versus local time for CHAMP during local time sectors at most latitudes. Figure 1 illustrates the
27 October to 2 November 2003. sampling of CHAMP in local time and latitude for the
period of 27 October to 2 November 2003, and Figure 2
shows the angle between the cross-track axis and geographic
associated with an X10 flare impacted the Earth. During this east for this time period. CHAMP’s high orbit inclination
storm, a southward IMF Bz component assured a severe thus allows for the study of zonal winds near equatorial
level of geomagnetic activity, during which time the plan- regions and meridional winds near the turning point of the
etary aP index remained saturated at 400 for 6 hours. Early orbit (near ±87) using the cross-track accelerometer axis.
on 31 October, activity began to decline, facilitated by a The combination of these capabilities provides an unprec-
northward IMF Bz component. Both CMEs had a transit edented opportunity to view the global density and wind
time of around 19 hours, making them among the fastest responses to the unusually severe solar and geomagnetic
recorded. In addition, this time span exhibited the highest disturbances discussed previously. The primary objectives
geomagnetic disturbances of the present solar cycle. of this paper are to describe the global density and wind
[3] The relationship between geomagnetic disturbances responses to these storms and to provide some interpreta-
and thermospheric composition, density, and winds has tion. In addition, we provide comparisons with current
been studied in depth for decades [Matuura, 1972; Prölss, empirical models of density and winds in order to identify
1980; Breig, 1987; Crowley, 1991]. The first discoveries of their capabilities and shortcomings.
this connection resulted from discrepancies between pre- [6] In the following section, we discuss the density
dicted and observed satellite ephemerides during periods of retrieval procedure, necessary force models, and semiempir-
geomagnetic activity. Early models for predicting thermo- ical models used for comparison. In section 3.1, we analyze
spheric density include the Jacchia total density models
[Jacchia, 1970] derived from satellite drag data. Zonal
winds were first studied by analyzing their average effect
on satellite orbits which was termed ‘‘superrotation’’ of the
atmosphere [King-Hele, 1964] and was estimated using
measurements of satellite inclination. More recent studies
of thermospheric conditions have utilized ground-based
incoherent scatter radar measurements of temperature and
in situ satellite measurements of total density and compo-
sition. The consolidation of these data is represented statis-
tically in the MSIS series of models [i.e., Hedin et al., 1977;
Hedin, 1983, 1987], and by the recent update to MSISe90
[Hedin, 1991], NRL-MSISe00 [Picone et al., 2002]. In a
similar fashion, a global empirical model of thermosphere
winds (HWM-93) was developed [Hedin et al., 1996] that
addresses a range of magnetic and solar conditions.
[4] The morphology of a geomagnetic storm is reviewed
in detail by Prölss [1980] and is related to changes in
composition, electric fields, neutral flow, and dynamo
effects all caused by an increase in energy input to the
high-latitude thermosphere-ionosphere system. In the Figure 2. The angle between the cross-track axis and
present paper, two elements indicative of these upper- geographic east for CHAMP during 27 October to
atmospheric disturbances are studied, namely in situ 2 November 2003.

2 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

density response to elevated levels of geomagnetic activity at where ~ aDRAG is the acceleration caused by drag, Ai is the
all latitudes and examine departures from the NRL-MSISe00 plate area, CDi is the coefficient of drag for the plate, m is
density model. In section 3.2, we analyze cross-track winds the satellite mass, ^ ni is the unit plate normal, r is the
with an emphasis on low-latitude response to elevated levels atmospheric density, and ~ vrel is the satellite velocity relative
of geomagnetic activity. Finally, a summary of our results to a corotating atmosphere. Typical in-track magnitudes for
and conclusions are given in section 4. this term vary from 2  107 to 12  107 ms2 over the
course of one orbit. This technique requires models to
2. Data and Models account for the unwanted forces that are measured along
with drag by the STAR accelerometer. For each of these
[7] CHAMP data was provided for our use by the models, orbit ephemeris files and quaternion files provided
CHAMP Information System and Data Center (ISDC). by the CHAMP ISDC are used to orient the satellite with
Acceleration, attitude, and orbit ephemeris data files were respect to inertial space, terrestrial coordinates, and the Sun.
used in the calculations pertaining to this study. The [10] The solar radiation pressure model also employs a
acceleration and attitude data is provided on a 10-s interval, 13-plate macromodel of the satellite. When the satellite is
which is processed from the original 1 Hz data to remove sunlit, attitude quaternions are used to calculate an angle
spikes caused by spacecraft maneuvers. Both the in-track between the satellite-Sun vector and the normal vector for
axis, used for density calculations, and the cross-track axis, each plate that is facing the Sun. The following equation can
used for wind calculations, are thought to be accurate to 3  then be used to sum the entire force of the satellite due to
109 ms2 [Reigber et al., 2002a]. The radial axis is not solar radiation [Luthcke et al., 1997]:
used in these studies because it is less sensitive by a factor
of ten and has had repeated trouble since launch.   
X
13
RAi cos finc;i crd;i  
[8] In order to have the most accurate density calculations, aSR ¼
~  2 þ crs;i cos finc;i ^ni
mc 3
a bias and scale factor must be applied to the raw acceleration i¼1

data. For this study, a least squares method was used to  
þ 1  crs;i ^s ; ð2Þ
estimate the bias factors for the in-track accelerometer axes.
With this method, the bias factors are estimated by using the
STAR accelerometer measurements as an orbit determination where ~ aSR is the acceleration caused by solar radiation
force model while letting the estimated orbit converge on the pressure, Ai is the plate area, c is the speed of light, crd,i is
Rapid Science Orbit ephemeris provided by the CHAMP the coefficient of diffusive reflectivity, crs,i is the coefficient
ISDC. The in-track a priori bias factor used for this time of specular reflectivity, m is the satellite mass, ^ ni is the unit
period was 2.925  106 ms2. The estimated values had a plate normal, finc,i is the angle of incidence of the Sun with
mean of 2.92534  106 ms2 and rms of 1.71  respect to the plate, R is the flux originating from the Sun,
108 ms2. We considered biases to be inaccurate when they and ^s is the unit satellite-Sun vector. The magnitude of solar
differed from the mean by more than 4.5  108 ms2. When flux is also multiplied by a ratio to account for shadowing
the numbers lie outside of this range, linear interpolation is when the satellite is in the umbra or penumbra shadow
used. An error of 1% in the scale factor translates to less than regions. The modifying ratio is equivalent to the percentage
1% of error in total density, while an error of 1% in the bias of the sun visible by the satellite. Finding the Sun-Earth
factor translates to an error of 2.5% in total density. For the vector and the appropriate flux acting on the satellite
cross-track axis, the bias factor was calculated assuming that requires up-to-date JPL solar and planetary ephemerides
on average for this time period, the measured cross-track (version DE-405). Typical in-track magnitudes for this term
acceleration agrees with the modeled accelerations which are on the order of 3  108 ms2 during the time period
assume a corotating atmosphere. For this time period, the studied in this work.
cross-track bias factor was determined to be 1.145978  [11] The albedo calculations require a latitudinally vary-
107 ms2. An error of 1% in the scale or bias factor ing model for short-wave radiation (in terms of albedo) and
translates to an error of about 5 – 10 ms1 in wind speed. long-wave radiation (in terms of emissivity) coming from
[9] On board the CHAMP satellite is the STAR acceler- the terrestrial sphere. The effect of the Earth is summed by
ometer, which measures the sum of all forces on the using discrete elements according to Knocke et al. [1988].
satellite’s surface. This measured quantity is comprised The following equation can be used to sum the effect of
mostly of the force imparted to the satellite by atmospheric each Earth element on each satellite plate:
drag, with lesser constituents such as solar and Earth
radiation pressure also contributing. All other nongravita-   
13 X
X Rj Ai cos finc;ij crd;i  
tional forces on the satellite are neglected in this study. aALB ¼
~  2 þ crs;i cos finc;ij ^ni
mc 3
Subtracting the modeled effects of radiation pressure, we are i¼1 j

left with the acceleration due to atmospheric drag. From  
þ 1  crs;i ^sj ; ð3Þ
here we arrive at a solution for atmospheric density which is
a modification of the traditional drag equation taking into
account a 13-plate macromodel of the satellite: where ~aALB is the acceleration caused by Earth radiation
pressure, Ai is the plate area, c is the speed of light, crd,i is
the coefficient of diffusive reflectivity, crs,i is the coefficient
of specular reflectivity, m is the satellite mass, ^ ni is the unit
r X13
aDRAG ¼ 
~ ½Ai CDi ð~
vrel  ^ni Þ~
vrel ; ð1Þ plate normal, finc,ij is the angle of incidence of the source
2m i¼1 with respect to the plate, Rj is the flux originating from the

3 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

source, and ^sj is the unit satellite-source vector. Both the 100 ms1. Meridional winds can skew density calculations
solar radiation pressure model and the albedo/infrared by 2.5% for every 100 ms1. Under quiet geomagnetic
models were adapted from the Geodyn II orbit determina- conditions, the error caused by horizontal winds rarely
tion software package (NASA Goddard Space Flight exceeds 15% near the poles and 8% away from the poles.
Center). Typical in-track magnitudes for this term are on However, under active conditions, the error can exceed 20%
the order of 5  1010 and 2  1010 ms2 for albedo and in the close vicinity of the poles.
infrared radiation pressure, respectively. [13] It is also possible to obtain an estimate of the neutral
[12] The final unknown in this process is the coefficient wind vector component in the direction of the cross-track
of drag for each plate of the CHAMP macromodel. Usually, axis of the STAR accelerometer. This calculation requires
these coefficients are solved for in terms of a density model the same steps as the density calculations. The following
using orbit determination software. However, for this appli- equation is now used as the drag equation:
cation, we desire that the coefficient of drag not have the
same bias that the corresponding thermospheric density rSTAR X
13
model possesses. Therefore a physics-based model for aDRAG ¼ 
~ fAi CDi ½ð~
vrel  ~
wÞ  ^
ni ð~
vrel  ~
wÞg; ð7Þ
2m i¼1
approximating the coefficient of drag for a flat plate was
used. This method is outlined by Cook [1965] and leads to
the following formula: where ~ aDRAG is the acceleration caused by drag, Ai is the
plate area, CDi is the coefficient of drag for the plate, m is
" sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

# the satellite mass, ^ni is the unit plate normal, rSTAR is the
2 Tw;i atmospheric density calculated from CHAMP/STAR, and
CDi ¼2 1þ 1 þ ai  1 cosðqi Þ ; ð4Þ
3 Ta ~
vrel is the satellite velocity with respect to a corotating
atmosphere. Here \vec{w} has been added as a vector
where CDi is the coefficient of drag, ai is the accommoda- representing wind in relation to an atmosphere that corotates
tion coefficient, Ta is the temperature of the atmosphere, Tw,i with the Earth. Only the cross-track axis is studied here for
is the temperature of the plate, and qi is the angle of incident the following two reasons. Neutral density and wind speed
gas flow with respect to the plate. Further, readmittance of cannot be separated from each other in calculations using
atmospheric molecules is assumed to be mainly diffusive the in-track accelerometer axis, and the radial accelerometer
because the temperature of the satellite is fairly cool axis does not provide measurements of sufficient accuracy.
(assumed to be 273.0 K). For macromodel plates that are Thus when solving equation (7) for wind speed, compo-
normal to the flow of atmospheric gases, this assumption nents of ~w in the in-track and radial directions are assumed
leads to an underestimation of CDi of about 2.5% for every to be zero. The density term in equation (7) is now the
100.0 K temperature increase in Tw,i. This assumption also neutral density derived from the in-track accelerometer axis.
leads to the following approximation for the accommoda- [14] The NRL-MSISe00 thermospheric density model
tion coefficient that appears in (4): [Picone et al., 2002] is used for comparisons to the
CHAMP/STAR density calculations. This model is an
3:6mi extension of the MSIS-83, -86, and -e90 versions [Hedin,
ai ¼ ; ð5Þ 1983, 1987, 1991]. All of these models are error weight
ð1 þ mi Þ2
nonlinear least squares fits of atmospheric composition data
where mi is the ratio of mass of the incident gas atom to the measured by mass spectrometers, and neutral temperature
mass of the surface atom. The force caused by impacting profiles measured by incoherent scatter radars, to empirical
and readmitting atoms and molecules can be represented by formulas. The NRL model also includes satellite drag data
the drag equation (1) in terms of the coefficient of drag, CDi. and calculations of anomalous oxygen in the atmosphere.
In this study, the coefficient of lift is assumed to be small The required inputs for the model are the satellite position,
enough to ignore. At its maximum, the force caused by lift day of the year, solar local time, F10.7 solar flux from the
is an order of magnitude less than that caused by solar previous day, mean F10.7 solar flux from 81 days centered
radiation. Estimating neutral density requires modeling on the current day, and a 57-hour history of the aP index.
surface forces other than drag, estimating instrument bias [15] The HWM-93 thermospheric wind model [Hedin et
and scale factors, and calculating the coefficient of drag. al., 1996], is used for comparisons to the CHAMP/STAR
After these steps are completed, (1) is solved for density cross-track wind calculations. The HWM is a error weight
using the in-track axis of the accelerometer denoted by nonlinear least squares fit to a truncated set of spherical
aSTAR  ^x, giving:
~ harmonics. It is primarily based on wind data obtained from
the AE-E and DE-2 satellites. Both satellites had highly
eccentric orbits. AE-E had an inclination of 20 and only
2mð~aSTAR  ~ aSR  ~ aALB Þ  ^x analyzed cross-track drift speeds, making zonal wind stud-
rSTAR ¼ P13 ; ð6Þ
i¼1 ½ A C ð~
v
i Di rel  ^
n Þð
i ~ vrel  ^xÞ ies difficult and high-latitude studies impossible. DE-2 had
a near-Polar orbit making it a better candidate for compar-
where rSTAR is the atmospheric density calculated from ison with CHAMP cross-track winds [Killeen and Roble,
CHAMP/STAR, and ^x is the unit vector of the in-track axis. 1988]. However, the DE-2 mission time span was too short
Normally, neutral wind velocity is considered to be the to capture any long-term variability. Inputs to the model
largest source of error in this calculation. From calculations include the satellite position, day of the year, solar local
using a simulated horizontal wind vector, we see that zonal time, F10.7 solar flux from the previous day, mean F10.7
winds can skew density calculations by 5% for every solar flux from 81 days centered on the current day, current

4 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

Figure 3. Latitude versus time, daytime response of CHAMP total mass densities at 410 km (top left),
the corresponding NRL-MSISe00 model densities sampled along the CHAMP orbit (top right), nighttime
response of CHAMP (bottom left), and corresponding NRL-MSISe00 model densities (bottom right).
Also shown are the Northern Polar Cap Index (top solid line in each panel), and 3-hourly Ap index
(bottom solid line in each panel).

3-hour aP index, and daily AP index. Lower-latitude winds latitude, local time is approximately 0720 near the North
are thought to be reproduced well by the model, while the Pole and 1920 near the South Pole. However, local time
structure of higher-latitude winds is somewhat lost by the sampling does not change substantially from the equator
simplicity of the spherical harmonic basis function. until CHAMP is above ±80 latitude. This complexity,
caused by the fact that CHAMP is not in a perfectly polar
orbit, is illustrated in Figure 1.
3. Results [17] During this period of elevated activity, there are two
3.1. Density Response time intervals in which the aP index becomes saturated at
[16] The daytime and nighttime global density responses 400, first for 3 hours early on 29 October (day 302) and then
at 410 km are illustrated in Figure 3. NRL-MSISe00 model again for 6 hours at the end of 31 October (day 304). In
response is also shown, sampled on the orbit of CHAMP, to between these two intervals, there is an active period of
draw comparisons and reveal any shortcomings of the about 9 hours where aP reaches 300. During the first surge
model. For each case, density response is shown in terms of geomagnetic activity, severe elevated levels of density
of latitude and time, with geomagnetic aP and northern can be seen confined to certain latitudes. On the dayside,
Polar Cap indices (Danish Meteorological Institute, http:// elevated density levels of order 300% (the highest of this
www.dmi.dk/projects/wdcc1/) indicated. Near the equator, time period) can be seen in high northern latitudes. How-
local solar time sampling is approximately 1320 and ever, this density enhancement seems to die out within an
0120 hours, respectively. When CHAMP is at maximum hour and a half, indicating that the first spike in geomag-

5 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

netic activity does not last long enough to have a global ing of the NRL-MSISe00 model is that it uses statistically
effect on density. On the nightside, density enhancements based empirical fits to represent density and hence tends to
last longer and can even be seen to travel toward the equator smear out density response features both spatially and
with time. During the second spike in which aP reaches 300, temporally. The effects of these issues can be seen in
density is more noticeably affected. On the dayside, a major Figure 3.
density disturbance of amplitude 200– 250% spans from [21] During times of elevated magnetic activity, daytime
85 latitude to the equator and lasts for upward of 9 hours. density measurements are higher than modeled density by
Much more activity can be seen in the southern hemisphere as much as a factor of 2. While modeled density seems to be
as well. Dayside density enhancements exhibit much less symmetric about the equator for day and nighttime density,
dispersion to lower latitudes in comparison to the nightside this is clearly not consistent with measured values, espe-
density disturbances. cially on the dayside.
[18] Several of the above response characteristics are [22] Note that just after 1200 UT on day 301, there is an
reminiscent of those revealed in accelerometer measure- enhancement in density of order 100% extending between
ments near 200 km, interpreted by Forbes et al. [1996] in 40 and +60 on the dayside but not on the nightside.
the context of the thermosphere modeling results of Fuller- Given that this is a period of very quiet geomagnetic
Rowell et al. [1996]. For instance the nighttime density activity, we believe that this is the response to a large solar
disturbances appear to more readily propagate to equatorial EUV flare that occurred near 1100 UT on day 301, as
latitudes at night, possibly facilitated by the predominant observed by the SEE instrument on the TIMED spacecraft
equatorward flow in both hemispheres at nighttime due to (T. Woods, private communication, 2003). A recent study of
the diurnal tides driven by in situ EUV heating [Forbes and these EUV events can be found in a paper by Tsurutani et al.
Garrett, 1976]. During daytime, the flow is predominantly [2005]. The possibility of an EUV flare response being
poleward, inhibiting equatorward extension of the density evident in CHAMP density data on day 301, and again
disturbances in both hemispheres. During both day and during a major solar EUV flare on day 308, is currently
night, horizontal advection and subsidence heating may be undergoing further investigation and will be reported on
playing a role in heating the low-latitude thermosphere. It is separately.
not understood why the daytime response in the summer [23] Equatorial density responses for the October through
hemisphere is less intense than that in the Northern Hemi- November time series can be seen in Figure 4, for daytime
sphere. Although this period is between equinox and and nighttime orbits. Both plots reveal a 3- to 6-hour delay
solstice, there is probably also a solstice-like net (diurnal between the onset of geomagnetic activity and equatorial
mean) meridional flow from the South Hemisphere to North density response. However, the amplitude of equatorial
Hemisphere which modulates the above dependence of the response to the first impulse of geomagnetic activity differs
density response at lower latitudes on time of day. This between dayside and nightside. During the night, disturban-
could also be caused by NRL-MSISe00 model inadequacy ces are free to travel equatorward, while the dayside
when attempting to normalize total density measurements to equatorial response is small. To an unknown extent, the
a height of 410 km. During this time period perigee of equatorial density enhancement may also be due to subsi-
CHAMP is close to the North Pole, which might explain an dence heating, facilitated by the nighttime equatorward flow
increased response in the Northern Hemisphere. connected with diurnal EUV heating.
[19] The nighttime response can also be viewed in the
context of traveling atmospheric disturbances (TADs). If we 3.2. Cross-Track Winds
assume that a nighttime TAD has a velocity of 750 ms1, [24] Studies of thermospheric winds are somewhat
successive orbits would observe movement between 32 and sparse, and in many cases, the methods used limit their
38 latitude, depending on whether the disturbance is comparability. For this reason, it is difficult to validate the
southbound or northbound. While there are many distur- measurements from CHAMP/STAR. Our estimates indi-
bances in this time period that have the potential of being a cate that CHAMP/STAR wind measurements are accurate
TAD, the large amount of latitudinal displacement between to about 60– 100 ms1. These estimates take into account
successive CHAMP orbits makes it difficult to identify a model error, bias factor error, accelerometer precision, and
TAD with any certainty. the error in the total density using a ‘‘propagation of
[20] The geomagnetic storm response using the NRL- uncertainty’’ statistical technique. Figure 5 illustrates the
MSISe00 model is driven by two parameters: a 57-hour behavior of the measured winds in the cross-track direction
history of the aP index and F10.7 solar flux measurements, for day and night orbits during days 300– 306 (27 October
adjusted to 1 AU. During geomagnetically quiet periods, to 2 November). Owing to the orbit constraints of CHAMP,
these two indices do fairly well at representing global the cross-track axis measures winds predominantly in the
density variations with a slight mean difference between east-west direction at middle to low latitudes (refer to
modeled values and measured values. This can be seen by Figure 2). However, above ±80 latitude, the cross-track
comparing prestorm modeled and measured density on 27 axis is oriented predominantly in the meridional direction,
and 28 October (day 300 and 301) in Figure 3. However, becoming fully north-south at the point of highest latitude
when this model is applied to storm time conditions, two for the orbit (±87). At the equator, red indicates eastward
problems arise: (1) Planetary indices do not contain enough winds for both day and night. Above ±80, red indicates
information to empirically emulate small-scale density northward winds for daytime orbits (top) and southward
structure, and (2) the aP index has a saturation point at winds for nighttime orbits (bottom). The HWM-93 wind
400 which precludes the model density from exhibiting the model has also been sampled on CHAMP’s orbit and the
same extrema as the measured densities. Another shortcom- vector winds projected into the cross-track direction to best

6 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

winds by the cross-track accelerometer. Near 1320 LT zonal


winds are consistently westward, on the order of 100 to
150 ms1 during times of low activity. The top panel of
Figure 6 shows a westward wind of 300 ms1 during the
period of highest activity at the beginning of day 304.
HWM-93 is consistent with CHAMP/STAR observations
during the quiet daytime. During active periods, however,
the model overestimates the observed wind enhancements
by about 50%, except during the aforementioned excursion
to 300 ms1 where the HWM-93 model yields 250 ms1.
[27] The bottom panel of Figure 6 compares nighttime
winds from CHAMP/STAR and HWM93. When geomag-
netic activity is low, the measured nightside winds are on
average only slightly positive ( 25 ms1), whereas HWM-
93 yields eastward winds of order 50 – 75 ms1. This
difference is within the estimated 60– 100 ms1 errors for
our wind measurements. While the dayside winds increase
in magnitude due to the increase in geomagnetic activity (cf.
top panel of Figure 6), nightside equatorial winds are seen
to reverse direction, reaching speeds as high as 275 ms1.
This shift in direction can also be seen to a lesser extent (0
to 50 ms1) in the HWM-93 model.
[28] The normal undisturbed behavior of the neutral
wind at equatorial latitudes according to the WATS
instrument on Dynamics Explorer [Wharton et al.,
1984; Wu et al., 1994] is westward during the day and
eastward at night, with maximum wind speeds of order
100 ms1. The majority of data from DE/WATS is
between 300– 400 km. This behavior is also reflected in
HWM-93, which is based in part on DE wind measure-
ments. However, analyses of equatorial DE/WATS winds
by Wu et al. [1994] and equatorial UARS/WINDII winds
near 250 km by Fejer et al. [2000] reveal rather small
differences between active and quiet geomagnetic con-
ditions. On the other hand, it is important to keep in
mind that the present events are unusually large in
Figure 4. Daytime (top) and nighttime (bottom) equatorial magnitude and that reversals of nighttime zonal winds
response of total density from CHAMP/STAR and NRL- have been observed under extremely active magnetic
MSISe00. Also shown is the 3-hour Ap index. conditions [Meriwether et al., 1986; Burnside et al.,
1991].
[29] The largest differences between CHAMP equatorial
illustrate similarities and differences. The HWM-93 results wind measurements and those consistent with HWM-93, are
are shown in the right panels of Figure 5. the unusually large westward accelerations at night occur-
[25] One of the most noticeable traits exhibited by the ring in conjunction with large geomagnetic disturbances.
thermospheric wind measurements is the combined effect of Changes in the zonal neutral wind speed are governed by
the longitude and universal time sampling of CHAMP, the zonal momentum equation:
consisting of a combination of the so-called ‘‘longitude-
UT’’ effect and the displacement between geographic and @Un
geomagnetic coordinates. These effects can be seen in the ¼  nin ðUn  Ui Þ þ pressure gradient term
@t
polar regions of both dayside and nightside orbits as an þ viscous term þ other terms; ð8Þ
apparent 24-hour periodicity, which is also reflected in the
HWM-93 model results. Owing to the complexity of the
cross-track sampling in the polar regions, it is not obvious where Un is the zonal neutral velocity, Ui is the zonal ion
how to connect our results with any specific neutral circu- velocity, and nin is the collision frequency of an ion with
lation pattern (i.e., two-cell), and we will not discuss the neutral gas [see Herrero et al., 1985]. Further, it is known
polar region wind results further. We simply note that the that in the equatorial region zonal winds tend to produce
overall magnitudes and patterns inferred from the cross- zonal ion drifts of similar magnitude and direction
track accelerations follow similar patterns to that of HMW- [Rishbeth, 1971; Fejer et al., 1985; Crain et al., 1993;
93, hence serving as a crude validation of the accelerometer Richmond et al., 1992]. Kamide and Matsushita [1981]
methodology for deriving neutral winds. speculated that during geomagnetically active periods, a
[26] Near the equator (see Figure 6), the high inclination low-latitude upward electric field is caused by penetration
of the CHAMP orbit permits a good measure of the zonal of high-latitude electric fields in the premidnight local time

7 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

Figure 5. Daytime response of CHAMP and HWM-93 cross-track wind speeds near 410 km (top). At
middle and low latitudes, red indicates eastward winds and near the poles red indicates northward winds.
Nighttime response of CHAMP and HWM-93 cross-track wind speed near 410 km (bottom). At middle
and low latitudes, red indicates eastward winds and near the poles red indicates southward winds.

region. The effect of this is a westward ion drift and hence namic interaction events that suggest further experimental
would lead to an increased westward neutral wind and modeling studies.
acceleration according to (8). This phenomenon has also
been observed in the postmidnight local time region at
Arecibo [Burnside et al., 1991], but in this case it was
4. Summary and Conclusions
deduced that an increase in ion-drag was responsible for a [30] We have shown that during times of extreme geo-
reversal of the normal eastward zonal wind. For the magnetic activity, thermosphere densities near 410 km
measurements presented here, it is not possible to unequi- exhibit enhancements of 200 – 300%. On the dayside, these
vocally determine if the observed westward surges at night enhancements are confined mostly to the Northern Hemi-
are due to changes in ion drifts or ion densities or both. sphere between the equator and 80 latitude. If not caused
Examination of GPS TEC measurements (International GPS by the height sampling of CHAMP, we cannot explain the
Service, http://igscb.jpl.nasa.gov) at low latitudes sampled origin of this latitudinal asymmetry in the density response.
near the CHAMP orbit do reveal, however, enhancements in During nighttime, density enhancements spread much more
the anomaly peaks during midday, and decreases in readily to equatorial latitudes, and we suggest that this is
equatorial-region TEC values at night (see Figure 7). Both facilitated by the nighttime equatorward flow that accom-
suggest equatorial penetration of electric fields and a panies the diurnal variation in EUV heating. Traveling
reduction in plasma collisions with neutrals. It is thus clear atmospheric disturbances would be expected to exist during
that the CHAMP measurements are revealing the neutral nighttime conditions based upon analyses of CHAMP
wind responses to significant plasma-neutral electrody- density data during the 15 – 24 April 2002 storm study

8 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

period (J. M. Forbes et al., manuscript in preparation, 2004).


However, owing to the sampling of CHAMP and the
complexity of the present storm interval, it is difficult to
unequivocally establish their presence or analyze their
movement.
[31] An apparent 24-hour periodicity appears both in the
high-latitude wind measurements and in HWM-93. We
interpret this as some combination of the so-called ‘‘longi-
tude-UT effect’’ and the displacement between geomagnetic
and geographic coordinates (i.e., the convection-driven
neutral winds tend to be ordered in the geomagnetic frame)
and moreover as a validation of the CHAMP winds.
Unfortunately, although at 87N and 87S the CHAMP
cross-track winds are in the meridional direction, these
measurements occur at dawn and dusk and do not sense
the significant antisunward flow that would be expected
from the convection driven circulation at these latitudes
under very disturbed conditions. Similarly, at ±60 latitude,
CHAMP is sampling near midday and midnight, far from Figure 7. Total electron content (TEC) sampled for
the dawn and dusk periods where significant convection- daytime CHAMP orbits (top) and nighttime CHAMP orbits
driven zonal winds might be expected to exist. Therefore (bottom).
and given the difficult geometry of the cross-track winds at
high latitudes and the associated ambiguities and uncertain-
ties, we have not attempted an extensive analysis of the high-latitude wind patterns. As a follow-up to this study, we
intend to obtain a better global view by comparing similar
measurements from the GRACE satellites which sampled
local times near 0400 and 1600 during this time period.
[32] Near the equator, measured zonal winds reveal
westward surges of order 100 – 200 ms1 in conjunction
with an increase in magnetic activity during both day and
night. The observed effect is shorter-lived (but of roughly
similar magnitude) than a similar feature contained in
HWM-93 under daytime conditions but exceeds the mag-
nitude of the nighttime HMW-93 response ( 100 ms1) by
about a factor of two. These data-model differences are not
surprising considering the large magnitudes of the geomag-
netic disturbances; however, it is also possible that this
difference is caused by an error in the cross-track bias
factor. It is not possible from our measurements to deter-
mine whether changes in plasma densities and/or zonal
plasma drifts driven by penetrating electric fields are the
root cause of the observed effects. We hope that our
experimental results will provide useful comparisons for
three-dimensional general circulation models of the thermo-
sphere and ionosphere that include self-consistent neutral
and plasma dynamics.
[33] For those interested in obtaining the CHAMP/STAR
winds and densities for further interpretation and
collaborative studies, please visit our web site (http://odo.
colorado.edu/ suttonek).

[34] Acknowledgments. The authors acknowledge the GeoFor-


schungsZentrum Potsdam (http://www.gfz-potsdam.de), who provided data
from the CHAMP satellite used for this study. The authors also thank Sean
Bruinsma for helpful discussions. This work was supported under grant
ATM-0208482 from the National Science Foundation as part of the
National Space Weather Program.
[35] Arthur Richmond thanks Sean Bruinsma and Douglas Drob for
their assistance in evaluating this paper.

References
Figure 6. Daytime (top) and nighttime (bottom) zonal
Berger, C., and F. Barlier (1981), Asymmetrical structure in the thermo-
wind speed at the geographic equator from CHAMP/STAR sphere during magnetic storms as deduced from the CACTUS acceler-
and HWM-93. Also shown is the 3-hour Ap index. ometer data, Adv. Space Res., 1(12), 231 – 235.

9 of 10
A09S40 SUTTON ET AL.: SEVERE GEOMAGNETIC STORMS OF 2003 A09S40

Breig, E. L. (1987), Thermospheric ion and neutral composition and chem- Killeen, T. L., and R. G. Roble (1988), Thermosphere dynamics: Contribu-
istry, Rev. Geophys., 25, 455 – 470. tions from the first 5 years of the Dynamics Explorer program, Rev.
Bruinsma, S., D. Tamagnan, and R. Biancale (2004), Atmospheric density Geophys., 26, 329 – 367.
derived from CHAMP/STAR accelerometer observations, Planet. Space King-Hele, D. G. (1964), The rotational speed of the upper atmosphere
Sci., 52, 297 – 312. determined from changes in satellite orbits, Planet. Space Sci., 12, 801.
Burnside, R. G., C. A. Tepley, M. P. Sulzer, T. J. Fuller-Rowell, D. G. Torr, Knocke, P. C., J. C. Ries, and B. D. Tapley (1988), Earth radiation pressure
and R. G. Roble (1991), The neutral thermosphere at Arecibo during effects on satellites, AIAA, 88 – 4292 – CP.
geomagnetic storms, J. Geophys. Res., 96, 1289 – 1301. Luthcke, S. B., J. A. Marshall, S. C. Rowton, K. E. Rachlin, and C. M. Cox
Cook, G. E. (1965), Satellite drag coefficients, Planet. Space Sci., 13, 929 – (1997), Enhanced radiative force modeling of the tracking and data relay
946. satellites, J. Astronaut. Sci., 45(3), 349 – 370.
Crain, D. J., R. A. Heelis, G. J. Bailey, and A. D. Richmond (1993), Low- Marcos, F. A., and J. M. Forbes (1985), Thermospheric winds from the
latitude plasma drifts from a simulation of the global atmospheric satellite electrostatic triaxial accelerometer system, J. Geophys. Res., 90,
dynamo, J. Geophys. Res., 98, 6039 – 6046. 6543 – 6552.
Crowley, G. (1991), Dynamics of the Earth’s thermosphere: A review, Rev. Matuura, N. (1972), Theoretical models of ionospheric storms, Space Sci.
Geophys., 29, 1143 – 1165. Rev., 13, 124 – 189.
Fejer, B. G., E. Kudeki, and D. T. Farley (1985), Equatorial F region zonal Meriwether, J. W., Jr., J. W. Moody, M. A. Biondi, and R. G. Roble
plasma drifts, J. Geophys. Res., 90, 12,249 – 12,255. (1986), Optical interferometric measurements of nighttime equatorial
Fejer, B. G., J. T. Emmert, G. G. Shepherd, and B. H. Solheim (2000), thermospheric winds at Arequipa, Peru, J. Geophys. Res., 91, 5557 –
Average daytime F region disturbance neutral winds measured by UARS, 5566.
Geophys. Res. Lett., 27(13), 1859 – 1862. Picone, J. M., A. E. Hedin, D. P. Drob, and A. C. Aikin (2002),
Forbes, J. M., and H. B. Garrett (1976), Solar diurnal tide in the thermo- NRLMSISE-00 empirical model of the atmosphere: Statistical compari-
sphere, J. Atmos. Sci., 33, 2226 – 2241. sons and scientific issues, J. Geophys. Res., 107(A12), 1468,
Forbes, J. M., R. Roble, and F. A. Marcos (1993), Magnetic activity de- doi:10.1029/2002JA009430.
pendence of high-latitude thermospheric winds and densities below Prölss, G. W. (1980), Magnetic storm associated perturbations of the upper
200 km, J. Geophys. Res., 98, 13,693 – 13,702. atmosphere: Recent results obtained by satellite-borne gas analyzers, Rev.
Forbes, J. M., R. Gonzalez, F. A. Marcos, D. Revelle, and H. Parish (1996), Geophys., 18(1), 183 – 202.
Magnetic storm response of lower thermosphere density, J. Geophys. Reigber, C., et al. (2002a), A high-quality global gravity field model from
Res., 101, 2313 – 2319. CHAMP GPS tracking data and accelerometry (EIGEN-1S), Geophys.
Fuller-Rowell, T. J., M. V. Codrescu, H. Rishbeth, R. J. Moffett, and Res. Lett., 29(14), 1692, doi:10.1029/2002GL015064.
S. Quegan (1996), On the seasonal response of the thermosphere and Reigber, C., H. Lühr, and P. Schwintzer (2002b), CHAMP mission status
ionosphere to geomagnetic storms, J. Geophys. Res., 101, 2343 – 2354. and perspectives, Eos Trans. AGU, 81(48), Fall Meet. Suppl., F307.
Hedin, A. E. (1983), A revised thermospheric model based on mass spec- Richmond, A. D., E. C. Ridley, and R. G. Roble (1992), A thermosphere/
trometer and incoherent scatter data: MSIS-83, J. Geophys. Res., 88, ionosphere general circulation model with coupled electrodynamics, Geo-
10,170. phys. Res. Lett., 19(6), 601 – 604.
Hedin, A. E. (1987), MSIS-86 thermospheric model, J. Geophys. Res., 92, Rishbeth, H. (1971), Polarization fields produced by winds in the equatorial
4649. F region, Planet. Space Sci., 19(3), 357 – 369.
Hedin, A. E. (1991), Extension of the MSIS thermosphere model in the Tsurutani, B. T., et al. (2005), The October 28, 2003 extreme EUV solar
middle and lower atmosphere, J. Geophys. Res., 96, 1159. flare and resultant extreme ionopheric effects, Geophys. Res. Lett., 32,
Hedin, A. E., et al. (1977), A global thermospheric model based on mass L03S09, doi:10.1029/2004GL021475.
spectrometer and incoherent scatter data MSIS: 1. N2 density and tem- Wharton, L. E., N. W. Spencer, and H. G. Mayr (1984), The Earth’s thermo-
perature, J. Geophys. Res., 82, 2139 – 2147. spheric superrotation from Dynamics Explorer 2, Geophys. Res. Lett., 11,
Hedin, A. E., et al. (1996), Emperical wind model for the upper, middle, 531 – 533.
and lower atmosphere, J. Atmos. Terr. Phys., 58, 1421 – 1447. Wu, Q., T. L. Killeen, and N. W. Spencer (1994), Dynamics Explorer-2
Herrero, F. A., H. G. Mayr, N. W. Spencer, A. E. Hedin, and B. G. Fejer observations of equatorial thermospheric winds and temperatures, J. Geo-
(1985), Interaction of zonal winds with the equatorial midnight pressure phys. Res., 99, 6277 – 6288.
bulge in the Earth’s thermosphere: Empirical check of momentum bal-
ance, Geophys. Res. Lett., 12, 491 – 494.
Jacchia, L. G. (1970), New static models of the thermosphere and exo- 
sphere with empirical temperature profiles, Spec. Rep. 313, Smithson. J. M. Forbes, R. S. Nerem, and E. K. Sutton, Department of Aerospace
Astrophys. Obs., Cambridge, Mass. Engineering Sciences, University of Colorado, UCB 429, Boulder, CO
Kamide, Y., and S. Matsushita (1981), Penetration of high-latitude electric 80309, USA. (forbes@colorado.edu; nerem@colorado.edu; eric.sutton@
fields into low latitudes, J. Atmos. Terr. Phys., 43, 411 – 425. colorado.edu)

10 of 10

You might also like