ArticlePDF Available

Large-scale structure of the fast solar wind

Authors:

Abstract and Figures

1] We present the results of a comprehensive study of the fast solar wind near solar minimum conditions using interplanetary scintillation (IPS) data taken with the EISCAT system in northern Scandinavia, and a recent extremely long baseline observation using both EISCAT and MERLIN systems. The results from IPS observations suggest that the fast wind inside 100 solar radii (R) can be represented by a two-mode model in some cases but this distinction is much less clear by in situ distances beyond 1 astronomical unit (215 R). Two distinct fast streams are seen in the extremely long baseline IPS observation; comparison of the IPS line of sight with a synoptic map of white light indicates the faster mode overlies the polar crown and the slower fast mode overlies an equatorial extension of the polar coronal hole.
Content may be subject to copyright.
Large-scale structure of the fast solar wind
M. M. Bisi,
1,3
R. A. Fallows,
1
A. R. Breen,
1
S. Rifai Habbal,
2
and R. A. Jones
1
Received 10 November 2006; revised 10 July 2006; accepted 13 March 2007; published 2 June 2007.
[1]We present the results of a comprehensive study of the fast solar wind near solar
minimum conditions using interplanetary scintillation (IPS) data taken with the EISCAT
system in northern Scandinavia, and a recent extremely long baseline observation using
both EISCAT and MERLIN systems. The results from IPS observations suggest that the
fast wind inside 100 solar radii (R
) can be represented by a two-mode model in some
cases but this distinction is much less clear by in situ distances beyond 1 astronomical unit
(215 R
). Two distinct fast streams are seen in the extremely long baseline IPS
observation; comparison of the IPS line of sight with a synoptic map of white light
indicates the faster mode overlies the polar crown and the slower fast mode overlies an
equatorial extension of the polar coronal hole.
Citation: Bisi, M. M., R. A. Fallows, A. R. Breen, S. Rifai Habbal, and R. A. Jones (2007), Large-scale structure of the fast solar
wind, J. Geophys. Res.,112 , A06101, doi:10.1029/2006JA012166.
1. Introduction
[2] Interplanetary scintillation (IPS) is the rapid variation
in the signal received by radio antennas at Earth from a
compact radio source arising from scattering by small-scale
(100 km) density variations in the solar wind. Measure-
ments of IPS allow the solar wind velocity to be inferred
over all heliographic latitudes and a wide range of helio-
centric distances [e.g., Dennison and Hewish, 1967]. When
two radio telescopes are used and the separation of the
raypaths in the plane of the sky from source to each
telescope is close to the radial direction centered at the
Sun, a high degree of correlation between the scintillation
patterns recorded at the two telescopes may be observed
[e.g., Armstrong and Coles, 1972; Coles, 1996]. The time
lag for maximum cross correlation can be used to estimate
the outflow speed of the irregularities producing the scin-
tillation. More sophisticated methods involving fitting the
results of a scattering model to the observed auto- and cross-
spectra may also be adopted [e.g., Coles et al., 1995; Coles,
1996; Klinglesmith, 1997].
[3] In weak scattering, the observed scintillation pattern
can be treated as arising from the sum of all the scattering
events along the raypath. If the raypath passes through two
or more streams of solar wind, their contributions are
independent and can be modelled separately provided that
their location along the raypath can be established. Studies
using the European Incoherent Scatter (EISCAT) IPS data
set have confirmed that the best results of fitting a weak
scattering model to the data [Coles, 1996; Klinglesmith,
1997] are obtained if regions of the raypath overlying dark
regions in white light observations are assumed to be
immersed in fast wind and slow streams are assumed to
overlie high intensity white light regions identified with
streamers [Coles, 1996]. Throughout these discussions we
define ‘‘polar crown flow’ as the central region of the polar
fast streams (as seen in interplanetary space above the large
polar coronal holes outside of solar maximum) and ‘‘flank
flow’’ as the region of fast flow between this crown outflow
and the equatorward boundary of the fast stream.
[4] Ulysses observations from the first polar pass (1994
1996) revealed the existence of a latitudinal gradient in the
fast wind speed across the polar regions, with the highest
velocities measured at the highest latitudes [Phillips et al.,
1995; Goldstein et al., 1996; Woch et al., 1997; McComas et
al., 2000; Habbal and Woo, 2001]. The boundary between
the fast and slow outflow measured by Ulysses was located
at about ±20°around the heliographic equator [Woch et al.,
1997].
[5] The most commonly accepted view of the fast wind
places its origin within polar coronal holes, at the centre of
the supergranular cells [Dupree et al., 1996] or the bound-
aries of supergranular cells [Hassler et al., 1999]. This view
implies a subsequent superradial expansion of the bound-
aries of the polar outflow so that the fast wind can occupy a
significant volume of the heliosphere, as observed by
Ulysses at solar minimum. However, using white light,
radio ranging and ultraviolet spectroscopic observations in
the inner corona, Woo and Habbal [1997] and Habbal and
Woo [2001] suggested that the fast wind originates in the
quiet Sun as well as in the polar coronal holes, and that
the outflow is predominantly radial. In the latter scenario,
the fast solar wind from the quiet Sun wind would be
expected to be marginally slower and slightly denser than
the outflow from the polar coronal hole.
[6] In this paper we use IPS data from EISCAT [Rishbeth
and Williams, 1985; Wannberg et al., 2002] taken during the
last solar minimum (1994 to 1997) supplemented by in situ
measurements from the solar wind observations over the
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, A06101, doi:10.1029/2006JA012166, 2007
1
Institute of Mathematical and Physical Sciences, University of Wales,
Aberystwyth, GB, UK.
2
Institute for Astronomy, University of Hawaii, USA.
3
Now at Center for Astrophysics and Space Sciences, University of
California, San Diego, La Jolla, California, USA.
Copyright 2007 by the American Geophysical Union.
0148-0227/07/2006JA012166$09.00
A06101 1of9
poles of the sun (SWOOPS) [Bame et al., 1992] instrument
on Ulysses, observations from Large Angle Spectroscopic
Coronagraph (LASCO) [Brueckner et al., 1995] and Extreme
Ultraviolet Imaging Telescope (EIT) [Delaboudiniere et al.,
1995] instruments on Solar and Heliospheric Observatory
(SOHO), as well as images from the Mauna Loa Mk III
coronagraph [Fisher et al., 1981], and the Soft X-Ray
Telescope (SXT) [Acton et al., 1989] instrument on Yohkoh
[Ogawara, 1987; Ogawara et al., 1991] to investigate
whether there is evidence of a two-mode structure within
the fast solar wind, as originally proposed by Woo and
Habbal [1997]. In carrying out this investigation, two main
questions were considered: First, is the solar wind best
represented by a two-mode fast wind; and second, is the
boundary between two modes of fast wind related to the
coronal hole boundaries seen in EUV and X ray. We first
describe the data (section 2), and the data analysis tools
used to infer flow speeds (section 3). The findings are
discussed in section 4, followed by concluding remarks in
section 5.
2. Observations
[7]Woo and Habbal [1997] suggested that a small step in
plasma number density seen in the Ulysses data indicated
the presence of two modes of fast wind, with the fastest (and
least dense) emerging radially from the polar coronal hole
and a second, slightly slower and denser stream emerging
from the quiet Sun. An EISCAT observation of the radio
source J0741+271 above the north polar coronal hole on
11 July 1995 appeared to be better represented not by fast
and slow streams but by a two-mode fast wind, prompting
a study into this question. In this study it was important to
consider only low-noise, high-latitude data from fast-wind
dominated observations of simple, isolated radio sources.
To reduce ground noise in the antenna sidelobes, only data
from observations at elevation angles of around 10°or
more were considered. The scattering model assumes that
all the scintillation comes from a single point source.
Structured or multiple sources in the beam will give rise
to superpositions of scintillation patterns which, at present,
cannot be fitted in a manner which is not misleading. The
sources used are all primary calibrators at 21-cm wave-
length for the ‘‘A’’ configuration of the VLA. Thus they
have no decorrelation over 35 km but they could still have
a small effect on the IPS, for which the Fresnel Radius at
an observing wavelength of 32 cm is approximately 90 km.
[8] For the scintillation to be strong enough for the signal
spectrum to be clearly resolved but not so strong as to
approach the strong scattering regime in the fast solar wind,
only observations for which the point of closest approach of
the raypath to the Sun lay between 15 R
and 120 R
were
considered. In the past, this has proven to be the best range
when considering mid- to high-latitude fast solar wind
observations at this frequency. This eliminates possible
distortion of the results by the effects of strong scattering.
[9] In order to detect small variations in the fast solar
wind structure, it is necessary to eliminate observations
containing a significant amount of slow wind. This was
done using a two-stage selection process. In the first stage
any observation in which less than 30°of the raypath
around the point of closest approach to the Sun lay above
the coronal hole was eliminated (A detailed description of
how sections of the raypath are identified as lying above a
coronal hole is given in the next section). This provided a
quick way of eliminating observations containing a large
proportion of slow solar wind. The list of observations was
then refined by calculating the proportion of scintillation
being generated in the region of the raypath which overlay
the coronal hole. Only observations in which more than
85% of the scintillation came from above the coronal hole
were considered for the main study. The slow wind was
assumed to scintillate 3.5 times more than the fast wind in
this calculation: this is likely to be an overestimate, as
EISCAT observations suggest that the density variations in
the slow wind are normally 2 3 times greater than those in
the fast wind over the distance range considered in this
study [Fallows et al., 2002], thus giving us considerable
confidence that we were indeed eliminating observations
which contained a significant proportion of slow wind.
[10] The majority of the IPS observations discussed in
this paper were made between 1994 and 1997 using the
EISCAT radio telescopes. The receiving frequency was on a
10 MHz bandwidth centered on 933.5 MHz from 1994 to
1995 and an 8 MHz bandwidth centered on 931.5 MHz
from 1996 to 1997. A single extremely long-baseline
observation combining data taken at 1420 MHz by EISCAT
and MERLIN telescopes in northern Scandinavia and the
UK, respectively [Bisi et al., 2005; Breen et al., 2006] is
discussed in section 4.3.
[11] The Ulysses data used in this paper are hourly
averaged SWOOPS ions radial velocity data taken during
the first polar pass of the Sun. The data were from the mid-
to high-latitude southern and northern polar passes where
the solar wind velocity did not dip below 600 km s
1
for
each hourly averaged data-point used, thus eliminating
interaction regions and periods of large variations in velocity
seen at the slow to fast wind transition in the Ulysses velocity
data.
3. Data Analysis
[12] A weak scattering interplanetary scintillation model
allowing for the presence of two solar wind streams in the
line of sight has been used for the last decade to analyze
EISCAT IPS observations [e.g., Coles, 1996; Klinglesmith,
1997; Canals, 2002]. The current analysis routines perform
all fitting to the auto- and cross-power spectra, with the
resulting correlation functions displayed for a more conve-
nient visualization.
[13] The power spectrum of the density variations in the
solar wind giving rise to IPS is described by a power law
with exponent a, cut off by an ‘‘inner scale’ q
i
corresponding
to the dissipation scale of the turbulence. Previous studies
[e.g., Yamauchi et al., 1996] have found ato vary between
2 and 4 and the inner scale in kilometers to be roughly equal
to the distance of the observation in solar radii, although it is
not very well determined. We adjust these parameters to
match the high-frequency tail of the autopower spectrum,
within these boundaries. The density variations are assumed
to be anisotropic and normally extended in the radial
direction. The model assumes that a fast stream will be
present in the central portion of the line of sight, flanked by
a slow stream at either end. The slow stream is given an
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
2of9
A06101
extra weighting to account for its greater density and hence
greater scattering power. Following the work of Little and
Ekers [1971], each stream can be modelled as having a mean
speed with random velocity components in the radial (dV
k
:
indicated by a skewed cross-correlation function) and
transverse (dV
?
: indicated by a reduced area under the
cross-correlation function) directions.
[14] This work required the model to be extended to
allow for a third stream in the line of sight. This had to
be done in such a way that the total number of free
parameters was not substantially increased. To achieve this,
the velocity of a second fast stream was modelled in place
of the dV
k
of the main fast stream. The second fast stream
was assumed to have the same remaining properties (dV
?
and axial ratio) as the main fast stream and to occupy a band
above the mainstreamer belt (Figure 1). The second fast
stream was given a small additional weighting of 1.5 to
account for the step in Ulysses’ density values noted by Wo o
and Habbal [1997]. This results in a model which can either
assume a single fast stream with a random radial component
or assume two discrete fast streams with the second
occupying a variable band above the streamer belt: an
increase of only one free parameter. The model will also
allow the width of the second fast stream to be different on
each side of the line of sight but this is only for the purpose
of fixing the fast/faster stream boundary to the coronal hole
boundary seen in EUV and X ray for a direct comparison.
No fit is performed involving both widths independently.
[15] The region of the IPS raypath overlying the polar
coronal hole, and hence assumed to be immersed in the fast
stream, is estimated by ballistically projecting the raypath
down onto a white light map constructed from either Mauna
Loa MkIII data or, when available, from LASCO observa-
tions. Determination of the region of the raypath overlying
the X-ray coronal hole is carried out in similar fashion using
synoptic maps constructed from SXT and EIT measure-
ments. In the analysis routines the portion of the raypath
above the coronal hole is described in terms of two Sun-
centered angles along the raypath from the point of closest
approach, q
in
(Earth side of the line of sight) and q
out
(source side of the line of sight). These two angles are
illustrated in Figure 1. The assumption of a fast stream
flanked by slow flow is normally a valid one since the
geometry of the line of sight between Earth and the radio
source usually ensures that the tail ends lie above the
equatorial streamer belt. Furthermore, since the level of
scattering giving rise to IPS falls as 1/R
4
, a fast stream in the
tail of the line of sight will not have any effect on the
measurements in comparison to the much denser slow
stream. However, slow flow in the tails of the line of sight
will still have an effect.
[16] Figure 2 shows a model fit to one of the observa-
tions, radio source J1120+143 on 8 September 1995, used
in this paper. The cross-correlation function shows a skewed
main peak with a kink at longer time lags indicating the
definite presence of a second stream. Mauna Loa MkIII
observations indicate that 1020°at the tail ends of the line
of sight lay above the streamer belt and so is assumed to be
in slow flow (Figure 3). The power law exponent awas set
to 3.0 (matching the high-frequency tail of the autopower
spectrum), the inner scale to 35 km (equal to the distance of
the point of closest approach) and the axial ratios of the fast
and slow streams to 2.3 and 1.5, respectively. Model fitting,
leaving the slow stream velocity set at 350 km s
1
, gave the
dominant fast stream a velocity of 872 km s
1
with a spread
of 193 km s
1
. However, if the slow velocity was also
fitted, it was found to rise to approximately 700 km s
1
,
fitting the kink in the cross-correlation function, but clearly
well above the velocity expected for solar wind flow above
the streamer belt. Hence for this fit, it was left fixed at 350
km s
1
, typical of slow wind speeds seen in IPS observa-
tions and by Ulysses.
[17] A fit using the new model is given for the 8 September
1995 observation in Figure 4. The model used the same
spectral parameters as before, but with two discrete fast
streams and a second stream width in place of the uniform
spread of velocities about the mean single fast stream.
Faster and fast velocities of 882 km s
1
and 621 km s
1
Figure 1. Diagram demonstrating the geometry of the new
model including two fast streams. The line of sight is
displayed as though it were mapped back down to the
corona. It should be noted that, although the second fast
stream width is fixed for both sides of the line of sight, each
tail end above the streamer belt may have a variable width.
Figure 2. The EISCAT IPS observation of J1120+143 on
8 September 1995, fitted with the two-stream weak scat-
tering model (dotted line). Following the method of analysis
described in the text, the power law exponent awas set to
3.0, the inner scale to 35 km and the axial ratios of the fast
and slow streams to 2.3 and 1.5, respectively. The dominant
fast stream has a velocity of 872 km s
1
with a random
radial component of 193 km s
1
. The slow stream has a
fixed velocity of 350 km s
1
.
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
3of9
A06101
respectively fitted best with a fast stream width of 21°. The
model clearly fits the kink in the cross-correlation function.
The quality of fit to the data using each model is best judged
by comparing the reduced-chi-square (c
n
2
) parameter, as
described in Reiff [1983], for the fits. In this example the
original two-stream model (two free parameters) gave c
n
2
=
1.69 and the new three-stream model (three free parameters)
gave c
n
2
= 0.99, indicatinga significantimprovementin the fit.
[18] The IPS data presented in this paper are therefore
analyzed in the following way. Only the mean fast stream
velocity and associated dV
k
are fitted; all other parameters,
including those of the slow stream, are chosen such that
they best represent the data (in terms of the overall shape of
the correlation functions) but are not fitted. The results of
these analyses were taken to represent the null-hypothesis
result, in which the fast solar wind was best represented by a
single stream with some finite variation in flow speed. Next,
in order to determine whether the observations could be
better represented using a two-mode fast wind, we used the
three-stream model to fit two discrete fast velocities and a
width in the line of sight for the second stream. All other
parameters were left as in the two-stream fit. The quality of
fit, as given by the c
n
2
, combined with the values obtained in
each fit are used to judge whether the observations are
best described by two discrete fast streams. The width
obtained for the second fast stream is then used to compare
the boundaries of both fast streams with coronal hole
boundaries seen in SXT and EIT synoptic maps. However,
since this fit restricts the width of the second fast stream to
be equal on each side of the line of sight, a further analysis
is performed in which the width is set according to the EUV
and X-ray coronal hole boundary and only the faster and
fast stream velocities are fitted.
4. Results
[19] Both EISCAT IPS observations and SWOOPS data
were analyzed as part of this investigation into the structure
of the fast solar wind.
4.1. EISCAT IPS Data
[20] The main case results are summarized in Table 1. Of
the eight observations that met all of the criteria for a clear
fast dominated flow, five of them (with c
n
2
values high-
lighted in bold in the table) were better fitted using the dual-
fast wind model. The faster of the two streams was the
stream at the higher latitude of the two streams. In all the
remaining cases, neither of the model fits could be preferred
over any other.
[21] Only one case showed a best fit when the boundary
between fast streams agreed with the coronal hole boundary
seen in X-ray/EUV. In all the remaining cases, the model
fitting indicated a boundary much lower in latitude than that
seen in X-ray/EUV. However, these results do not indicate
any strong or systematic relationship between the fast
Figure 3. A white light map of the east limb of the Sun, Carrington rotations 1889 1900 centered on
8 September 1995, constructed using data taken from the Mauna Loa MkIII Coronagraph at 1.7 R
with
the IPS raypath mapped onto it and both SXT-5°latitude and MkIII coronal hole boundaries circled as
well as the P-point, the point of closest approach of the IPS raypath to the Sun in the plane of the sky.
Figure 4. The EISCAT IPS observation of J1120+143 on
8 September 1995, fitted with the three-stream weak
scattering model (dotted line). The spectral and slow stream
parameters are set as in Figure 2. The faster and fast
velocities are 895 and 661 km s
1
, respectively, with a fast
stream width of 18°.
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
4of9
A06101
stream boundaries inferred from the IPS analyses and those
obtained from EUV/X-ray Carrington maps. We therefore
consider it important to compare these results with the one
and only available set of high latitude in situ solar wind
data recorded, the SWOOPS measurements made by the
Ulysses spacecraft while making its first polar pass at solar
minimum.
4.2. Ulysses Swoops in Situ Data
[22] The Ulysses SWOOPS ions hourly averaged data
were ‘‘binned’ (averaged further) for approximately every
100 data points. In order to restrict the study to fast wind, a
low-latitude cut-off point was established at the lowest
latitude above which the solar wind velocity never dipped
below 600 km s
1
in any of the original hourly averaged
data points. The data were split into four different sections
corresponding to the increasing and decreasing latitude
parts of the northern and southern polar passes. Each section
was then fitted with polynomials of orders 1 to 3 via a least
squares fitting method. A bimodal linear fit (i.e., straight
line with a change in gradient) was also performed with the
breakpoint between the two linear fits set to a range of
latitudes from 75°N/S to 10°from the low-latitude cut-off.
[23] Using the c
n
2
value as a measure of the goodness of
fit, the data appeared to be best represented by the bilinear
fit in three out of the four segments and a linear fit in one
out of the four segments as shown in Tables 2 and 3, and in
Figure 5. It should be noted that in those cases where a
bilinear fit gave the lowest c
n
2
value closest to 1 in Table 2,
the improvement of fit gained was generally very small. It
should also be noted that in the increasing latitude northern
pole case, the c
n
2
value is always overfitted which is why the
straight-line linear fit is taken as the most appropriate in this
case. Table 3 gives details of the best least squares fit in
each case, along with the latitudes of the polar coronal holes
as seen in SXT and EIT.
[24] There was no clear relationship between the latitude
of the inflexion point of the bilinear fits and the latitude of
the coronal hole boundary seen in X-ray or EUV emission
during the same Carrington rotation. The only clearly
significant result indicated a single fast wind with higher
velocities at higher latitudes, which is in good agreement
with the results of McComas et al. [2000].
[25] Figure 6 shows a rough comparison of the two-mode
fast IPS best-fitted data and ion SWOOPS data, bearing in
mind that the observations were taken at different times and
so cannot be taken as a direct comparison. However, they
can be used to get the general picture of the two different
data sets. To save overcrowding on the plot, only the most
significantly fitted Ulysses data from each hemisphere are
used. The fast velocities and their associated spread (dV
k
)
from the traditional two-stream model fits are used as an
illustration of the range of velocities given by the IPS
modeling. dV
k
was not modelled in the two-mode fast wind
model fitting and so no spread in velocity can be inferred
from these results. Instead, they are used to illustrate the
range of latitudes covered by each fast stream, as
determined by the fast/slow wind boundaries and the fast/
faster wind boundaries.
4.3. Extremely Long Baseline Observation
[26] Recent developments in the IPS observing technique
have included the use of much more widely separated
antennas with separations up to 2000 km in the radial
baseline, carried out with EISCAT and the Multi-Element
Radio Linked Interferometer Network (MERLIN) based in
Table 1. Summary of IPS Results
Date Source V
fast
dV
k
V
mid
Wdth
in
Wdth
out
rmsVV
slow
q
in
q
out
wtslow AR aq
i
c
n2
14 July 1995 J0741+312 754 444 80 300 65°70 8 2.0 3.0 36 2.96
842 443 31°31°1.72
886 456 35°40°2.07
6 September 1995 J1120+143 811 211 0 450 70°60°3 2.3 3.0 35 5.83
844 548 23°23°3.57
931 755 50°40°6.03
7 September 1995 J1120+143 872 210 0 450 70°60°2 2.0 3.0 35 3.70
895 708 34°34°2.90
929 794 45°35°3.15
8 September 1995 J1120+143 872 193 0 350 70°60°2 2.3 3.0 35 1.69
882 621 21°21°0.99
951 831 50°35°1.70
29 May 1996 J0336+323 827 146 0 350 75°75°8 1.5 3.4 57 1.68
842 738 42°42°1.47
901 827 65°45°1.64
8 June 1996 J0521+166 820 126 0 350 65°80°8 1.6 3.2 26 6.59
867 770 51°51°6.22
856 826 55°30°6.49
9 June 1996 J0521+166 774 215 0 350 70°80°4 1.5 3.1 25 3.90
875 708 54°54°3.05
878 706 55°55°2.99
17 August 1996 J0954+177 827 408 0 450 60°60°10 3.5 2.5 8 1.14
934 609 26°26°1.04
968 644 35°35°1.15
Table 2. Summary of SWOOPS Ions Analysis
a
Type/Order of fit
South
increasing
lat.
South
decreasing
lat.
North
increasing
lat.
North
decreasing
lat.
Straight line 2.77866 1.00748 0.48211 1.13374
Two-stage gradient 2.36277 1.00580 0.46569 1.06449
X
2
parabola 2.42265 1.02292 0.47697 1.10716
X
3
parabola 2.39792 1.05332 0.49470 1.11038
a
The lowest c
n2
value closest to 1 are in bold.
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
5of9
A06101
the United Kingdom (this experiment is described further by
Breen et al. [2006]). The difference between time lags for
solar wind speed increases as the antenna separation gets
larger. An observation of J0319+415 from 12 May 2004
produced a correlation function that showed two clear fast
wind peaks [Bisi et al., 2005; Breen et al., 2006].
[27] This observation strongly suggested the presence of
two distinct fast components which could be interpreted in
two ways:
[28] 1. As a region of faster flow from above the polar
coronal hole at the crown and fast flow above the quiet Sun; or
Table 3. Summary of SWOOPS Ions Best Fit Analysis
Variable South increasing lat. South decreasing lat. North increasing lat. North decreasing lat.
Latitude of change in gradient (52.95 +2.58/1.50)°(72.48 +0.48/0.93)°Straight line (+31.94 +1.95/1.04)°
Date of change in gradient 16 February 1994, 46.60 30 October 1994, 302.23 N/A 11 July 1996, 192.48
Carrington rotation during
change in gradient
1879 1888 N/A 1911
Days of year for the
Carrington rotation
1994-37.96-65.29 1994-283.24-310.53 N/A 1996-180.70-207.90
Latitude of SXT coronal hole 60 ± 10°50 ± 10°50 ± 10°50 ± 10°
Latitude of EIT coronal hole Not available Not available Not available 60 ± 10°
c
n2
of fit 2.362770 1.005800 0.482111 1.064490
Nu 94 26 27 89
Significance Level 0.001 0.450 >0.990 0.300
Summary No clear fit Weak evidence for a
bi-modal fast wind
Strong evidence for a
single fast wind
Weak evidence for a
bi-modal fast wind
Figure 5. Plots of averaged Ulysses ions radial velocity for the first polar pass. The two lines represent
the ‘‘best fit’ determined by the c
n2
fit to the binned data showing that a change in gradient of velocity
with latitude fits the data best. The arrow indicates the inflexion point between the two lines. The latitude
and c
n2
fit values can be seen in Table 3. (a) South polar pass increasing in latitude; July 1993 to
September 1994. (b) South polar pass decreasing in latitude; September 1994 to January 1995. (c) North
polar pass increasing in latitude; March 1995 to July 1995. (d) North polar pass decreasing in latitude;
August 1995 to August 1996.
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
6of9
A06101
[29] 2. As faster flow from above the whole of the white
light coronal hole and fast flow above the equatorward
extension of the polar coronal hole.
[30] The outflow from equatorward extensions of coronal
holes is generally accepted to be slower than that from large
polar holes [e.g., Kojima et al., 2004]. Figure 7 shows the
line of sight projected back along the Parker spiral (assum-
ing radial outflow) and down to the solar corona at 2.5 R
and overlaying a synoptic map of white light. The geometry
of the observation does suggest that the Earthward side of
Figure 6. Figure showing the IPS fitted velocity values as a rough comparison to the values of the two
most significantly fitted Ulysses SWOOPS ions radial velocity binned data with latitude (south
decreasing latitude and north increasing latitude). Two-stream modelled velocities (blue) are included
simply as a measure of the range of velocities given by the IPS modeling, using half dV
k
as the error; the
latitude given is that of the point of closest approach. The two-mode fast wind modelled fast (red) and
faster (green) velocities used are from the best fit from each observation using errors quoted from the
least squares fitting routine. The latitude ranges given are derived from the average latitudes of the fast/
faster wind boundaries and the fast/slow wind boundaries.
Figure 7. Synoptic map of LASCO east limb white light data from Carrington rotation 2016, overlayed
by the IPS line of sight to J0319+415. The arrows indicate the positions of the fast/slow wind boundaries
(as inferred from this image) and the fast/faster wind boundary as fitted using the two-mode fast wind
model.
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
7of9
A06101
the line of sight lies above an equatorward extension of the
polar coronal hole. The transition between flow above a
large polar hole and that above an equatorward extension
would be fairly abrupt and lead to the appearance of two
discrete fast components of flow.
[31] To account for this scenario, a further modification
was made to the analysis program to allow the width of the
fast stream to be fitted on only one side of the line of sight.
The IPS data were analyzed using the two-mode fast wind
model (Figure 8) and the results are unambiguous: It is only
possible to reproduce the shape of the cross-correlation
function if the fast stream occupies part of the line of sight
over the equatorward extension of the coronal hole. No
other scenario would reproduce the correct shape. The
stream boundaries in the IPS line of sight, as found in this
analysis, are also marked in Figure 7.
[32] This result provides convincing evidence of a greater
degree of nonuniformity in the fast solar wind than is seen
in the Ulysses|SWOOPS data.
5. Discussion
[33] The results from IPS and in situ measurements
reported in this paper provide evidence for a change in
character of the solar wind between the polar crown and the
equatorward flanks of the fast wind, but IPS results suggest
a two-mode fast solar wind more strongly than do the in situ
data.
[34] The polar fast wind is faster and appears to vary little
with latitude whereas the wind above the lower latitude
regions of the white light coronal hole shows a more
marked decrease in velocity with decreasing latitude. These
results are generally consistent with the Woo and Habbal
proposals. However, the Ulysses results suggest a more
continuous distribution of solar wind speeds rather than two
distinct modes.
[35] Although the majority of the IPS results indicate that
a two-mode fast wind model is a better fit to the measure-
ments, the c
n
2
values are high in many cases indicating that
the model still does not adequately describe the data. A
possible explanation for this is the steady variation in
velocity with latitude of the flanking fast stream, as seen
in the Ulysses data at lower latitudes, which has not been
included in the two-mode fast wind IPS model. The fact that
some observations do not fit significantly better with the
two-mode fast wind model also suggests this explanation.
[36] The boundary between fast and faster streams in the
IPS lines of sight modelled in the IPS analysis showed no
correlation with the coronal hole and quiet Sun boundary
seen in EUV and X-ray measurements. This suggests that
either there is no distinction between the fast wind emanat-
ing from the quiet Sun and the EUV/X-ray coronal hole, or
that the individual fast wind components have interacted to
a large enough extent for the boundary to be significantly
blurred by IPS distances.
[37] The extremely long baseline observation of 12 May
2004 does suggest two distinct modes of fast wind. How-
ever, in this case, the raypath extended from above the polar
hole into a region above an equatorward extension of the
northern hole. Analysis unambiguously indicated that the
slower fast stream lay above the equatorward extension of
the coronal hole. We therefore propose that the observation
is dominated primarily by high-latitude, low-density flow
from across the raypath near the point of closest approach,
and secondly, by slightly slower higher density flow above
the equatorward extension of the coronal hole.
6. Conclusions
[38] The results presented in this paper provide a useful
test of the alternative models of fast wind internal structure.
They reveal a difference in character between the flow
found in interplanetary space above the polar crown and
above the flanks of the white light coronal hole, though
probably not a sudden transition from one type of flow to
another. Within the fast solar wind the latitudinal gradient in
velocity in the polar crown stream is considerably shallower
than that seen in the equatorward flanks of the fast stream.
However, the Ulysses data do not provide significant
evidence of a bimodal fast wind, though it could be the
case that there is an inhomogeneity in the turbulence that
damps out between IPS distances (typically 50 R
) and
the typical distance of Ulysses (over 10 times that). It can
only suggest that there is a difference in character between
the polar stream and the flanks of the fast wind.
[39] We suggest that the flow observed above the centers
of the polar coronal holes should be regarded as the
archetypal fast wind. The change in latitudinal gradient in
velocity observed in interplanetary space at latitudes near
to those of the boundaries of the X-ray coronal holes may
indicate an internal boundary within the fast wind or may
simply be the signature of a more gradual change in solar
wind properties. The presence of two distinct fast velocities
in the EISCAT-MERLIN result strongly suggests a differ-
ence in character between the high-latitude fast wind
(found above the polar coronal hole) and the flow above
the equatorward extension of the polar hole. It is also likely
that this difference applies to flow above the centers of
polar coronal holes and regions near their equatorward
boundaries.
Figure 8. The EISCAT/MERLIN observation of J0319+415
on 12 May 2004, fitted with the two-mode fast wind
weak scattering model. Parameters are: V
faster
766 km s
1
;
V
fast
640 km s
1
; axial ratio 2.0; q
in
55°;q
out
55°;V
slow
350 km s
1
; weighting 5.3; a3.0; inner scale 85.0; fast
stream width 40°(Earth end of the line of sight only). The
c
n2
is 0.51.
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
8of9
A06101
[40]Acknowledgments. We would like to thank the directors and
staff of EISCAT and MERLIN for the IPS data used in this paper and for the
supply of equipment while using these facilities. EISCAT is supported by
the scientific research councils of Finland, France, Germany, Japan, Nor-
way, Sweden, and the United Kingdom. The MERLIN system is supported
by PPARC, as were several of the authors of this paper (RAF, MMB, and
RAJ) during the time this work was carried out. The programs used to
analyze the IPS data are based on routines developed by W. A. Coles and
B. J. Rickett at the University of California, San Diego and we are
extremely grateful not only for the programs but also for much valuable
advice. LASCO data are used courtesy of the LASCO consortium and were
made available by the EIT consortium. We would also like to thank D. J.
McComas and the SWOOPS team for the solar wind plasma data,
A. Lecinski for the MkIII coronagraph white light data, and the Yohkoh
consortium for the SXT data.
[41]Amitava Bhattacharjee thanks W. A. Coles and another reviewer
for their assistance in evaluating this paper.
References
Acton, L. W., M. Bruner, W. Brown, J. Lemen, T. Hirayama, S. Tsuneta,
T. Watanabe, and Y. Ogawara (1989), The solar-A soft X-ray telescope
experiment, Adv. Space Res.,8(11), 93 99.
Armstrong, J. W., and W. A. Coles (1972), Analysis of three-station inter-
planetary scintillation data, J. Geophys. Res.,77, 4602 4610.
Bame, S. J., D. J. McComas, B. L. Barraclough, J. L. Phillips, K. J. Sofaly,
J. C. Chavez, B. E. Goldstein, and R. K. Sakurai (1992), The Ulysses
solar wind plasma experiment, Astron. Astrophys. Suppl.,92(2), 237.
Bisi, M. M., A. R. Breen, R. A. Fallows, P. Thomasson, R. A. Jones, and
G. Wannberg (2005), Combined EISCAT/ESR/MERLIN interplanetary
scintillation observations of the solar wind, in Solar Wind 11/SOHO 16,
Connecting Sun and Heliosphere, edited by B. Fleck, T. H. Zurbuchen,
and H. Lacoste, Eur. Space Agency, ESA SP-592, 593 596.
Breen, A. R., R. A. Fallows, M. M. Bisi, P. Thomasson, C. A. Jordan,
G. Wannberg, and R. A. Jones (2006), Extremely long-baseline interpla-
netary scintillation measurements of solar wind velocity, J. Geophys.
Res.,111, A08104, doi:10.1029/2005JA011485.
Brueckner, G. E., et al. (1995), The large angle spectroscopic coronagraph
(LASCO), Sol. Phys.,162, 357 402.
Canals, A. (2002), Interplanetary scintillation studies of the solar wind
during the rising phase of the solar cycle, Ph.D. thesis, The University
of Wales, Aberystwyth.
Coles, W. A. (1996), A bimodal model of the solar wind speed, Astrophys.
Space Sci.,243(1), 87 96.
Coles, W. A., R. R. Grall, T. Klinglesmith, and G. Bourgois (1995), Solar
cycle changes in the level of compressive mircoturbulance near the sun,
Geophys. Res., 100.
Delaboudiniere, J. P., et al. (1995), EIT: Extreme-Ultraviolet Imaging Tele-
scope for the SOHO Mission, Sol. Phys.,162, 291 312.
Dennison, P. A., and A. Hewish (1967), The solar wind outside the plane of
the ecliptic, Nature,213, 343 346.
Dupree, A. K., M. J. Penn, and H. P. Jones (1996), He i 10830 Angstrom
wing asymmetry in polar coronal holes: Evidence for radial outflows,
Astrophys. J.,467, L121+.
Fallows, R. A., P. J. S. Williams, and A. R. Breen (2002), EISCAT mea-
surements of solar wind velocity and the associated level of interplanetary
scintillation, Ann. Geophys.,20, 1279 1289.
Fisher, R. R., R. H. Lee, R. M. MacQueen, and A. I. Poland (1981), New
Mauna Loa coronagraph systems, Appl. Opt.,20, 1094 1101.
Goldstein, B. E., M. Neugebauer, J. L. Phillips, S. Bame, J. T. Gosling,
D. McComas, Y.-M. Wang, N. R. Sheeley, and S. T. Suess (1996),
ULYSSES plasma parameters: Latitudinal, radial, and temporal varia-
tions, Astron. Astrophys.,316, 296 303.
Habbal, S. R., and R. Woo (2001), Connecting the Sun and the solar wind:
Comparison of the latitudinal profiles of coronal and Ulysses measure-
ments of the fast wind, Astrophys. J.,549, L253 L256.
Hassler, D. M., I. E. Dammasch, P. Lemaire, P. Brekke, W. Curdt, H. E.
Mason, J. C. Vial, and K. Wilhelm (1999), Solar wind outflow and the
chromospheric magnetic network, Science,283, 810.
Klinglesmith, M. (1997), The polar solar wind from 2.5 to 40 solar radii:
Results of intensity scintillation measurements, Ph.D. thesis, University
of California, San Diego.
Kojima, M., A. R. Breen, K. Fujiki, K. Hayashi, T. Ohmi, and M. Tokumaru
(2004), Fast solar wind after the rapid acceleration, J. Geophys. Res.,
109(A18), A04103, doi:10.1029/2003JA010247.
Little, L. T., and R. D. Ekers (1971), A method for analysing drifting
random patterns in astronomy and geophysics, Astron. Astrophys.,10,
306 309.
McComas, D. J., B. L. Barraclough, H. O. Funsten, J. T. Gosling,
E. Santiago-Munoz, R. M. Skoug, B. E. Goldstein, M. Neugebauer,
P. Riley, and A. Balogh (2000), Solar wind observations over Ulysses’
first polar orbit, J. Geophys. Res.,105(A5), 10,419 10,433.
Ogawara, Y. (1987), The SOLAR-A mission, Sol. Phys.,113 , 361 370.
Ogawara, Y., T. Takano, T. Kato, T. Kosugi, S. Tsunta, T. Watanabe,
I. Kondo, and Y. Uchida (1991), The Solar-A Mission: An overview,
Solar Physics, 136, 1 16, Springer, New York.
Phillips, J. L., S. J. Bame, W. C. Feldman, B. E. Goldstein, J. T. Gosling,
C. M. Hammond, D. J. McComas, M. Neugebauer, E. E. Scime, and
S. T. Suess (1995), ULYSSES solar wind plasma observations at high
southerly latitudes, Science,268, 1030±.
Reiff, P. H. (1983), The use and misuse of statistical analyses, in ASSL Vol
104: Solar-Terrestrial Physics: Principles and Theoretical Foundations,
p. 493, Springer, New York.
Rishbeth, H., and P. J. S. Williams (1985), Ionospheric radar: The system
and its early results, Mon. Not. R. Astron. Soc.,26, 478 512.
Wannberg, G., L.-G. Vanhainen, A. Westman, A. R. Breen, and P. J. S.
Williams (2002), The new 1420 MHz dual polarisation interplanetary
scintillation (IPS) facility at EISCAT, in Conference Proceedings, Union
of Radio Scientists (URSI).
Woch, J., W. I. Axford, U. Mall, B. Wilken, S. Livi, J. Geiss, G. Gloeckler,
and R. J. Forsyth (1997), SWICS/ULYSSES observations: The three-
dimensional structure of the heliosphere during solar minimum, Geophys.
Res. Lett.,24, 2885 2888.
Woo, R., and S. R. Habbal (1997), Extension of coronal structure into
interplanetary space, Geophys. Res. Lett.,24(10), 1159 1162.
Yamauchi, Y., M. Kojima, M. Tokumaru, H. Misawa, H. Mori, T. Tanaka,
H. Takaba, T. Kondo, and P. K. Manoharan (1996), Micro-turbulence in
the solar wind at 5–76 rs observed with interplanetary scintillation,
J. Geomagn. Geoelectr.,48, 1201 1217.
M. M. Bisi, Center for Astrophysics and Space Sciences, University of
California, San Diego, 9500 Gilman Dr., La Jolla, CA 92093, USA.
(mmbisi@ucsd.edu)
A. R. Breen, R. A. Fallows, and R. A. Jones, Institute of Mathematical
and Physical Sciences, University of Wales, Physical Sciences Building,
Penglais Hill, Aberystwyth, Ceredigion SY23 3BZ, UK. (raf@aber.ac.uk)
S. Rifai Habbal, Institute for Astronomy, University of Hawaii, USA.
A06101 BISI ET AL.: LARGE-SCALE STRUCTURE OF THE FAST SOLAR WIND
9of9
A06101
... When baselines are sufficiently long enough between two different sites simultaneously observing the same radio source, there is a possibility to resolve more than one velocity/structure crossing the LOS. Bisi et al. (2007a) have measured up to three distinct velocity streams; this signature is observed as different peaks in the CCF. This is not something which can be gleaned easily from a single power spectrum alone. ...
... This is not something which can be gleaned easily from a single power spectrum alone. However, there are some small indications (which require a much deeper analyses) that certain features in a single-power spectrum can relate to the fundamental microscale features on the plasma flowing out from the Sun, which can be tied to the well-known features of a CCF (e.g., Bisi, 2016a;Bisi et al., 2007a;Coles, 1996;Jones et al., 2007;Klinglesmith et al., 1996, and references therein). One example here is that typically the passage of a CME can be related to a negative-correlation lobe at (or near) zero time lag. ...
Article
Full-text available
Interplanetary scintillation (IPS) manifests itself as a variation in the radio signal received from a distant, compact radio source on the sky. The intensity fluctuations of the radio waves are caused by density inhomogeneities in the outflowing solar plasma across the heliosphere. IPS allows us to infer solar wind speed and density variations along the line of sight. There are two types of techniques in the literature to infer solar wind speed using IPS data sets: single‐site analysis (SSA), where the power spectra from single time series are analyzed to obtain solar wind parameters, and multisite analysis, where the cross‐correlation function of data from two or more widely separated sites is used. The selection of the analysis technique depends on the number of sites available to each IPS system. In order to combine and complement solar wind speed determinations from different instruments, it is important to validate results and methodologies of the two techniques. In this paper, we analyzed previously well‐studied European Incoherent SCATter (EISCAT) and Multi‐Element Radio‐Linked Interferometer Network (MERLIN) observations of IPS with well‐known results from the cross‐correlation function methodology. We applied the SSA technique to each of the individual EISCAT and MERLIN IPS power spectra. This work shows the capabilities of the SSA to describe complex events and seeks to obtain improved parameter fits using the SSA methodology.
... Observing stations with radio source tracking capabilities can make multiple passes during a single day, or dwell on individual sources for a longer period of time, and simultaneous observation from such system(s) with stations several hundred or more kilometres apart are capable of getting much more detail from single observations. Such observations enable multiple solar wind streams to be detected crossing the observing station to radio source lines of sight, such as the "fast and faster" solar wind streams detected by the Ulysses spacecraft and observed in IPS using the combined European Incoherent Scatter (EISCAT) and Multi-Element Radio-Linked Interferometer Network (MERLIN) systems (Bisi et al., 2007). Furthermore, longer-duration observations are possible which enable changes in solar wind structure (e.g. ...
Preprint
Full-text available
Observations of interplanetary scintillation (IPS - the scintillation of compact radio sources due to density variations in the solar wind) enable the velocity of the solar wind to be determined, and its bulk density to be estimated, throughout the inner heliosphere. A series of observations using the Low Frequency Array (LOFAR - a radio telescope centred on the Netherlands with stations across Europe) were undertaken using this technique to observe the passage of an ultra-fast CME which launched from the Sun following the X-class flare of 10 September 2017. LOFAR observed the strong radio source 3C147 at an elongation of 82 degrees from the Sun over a period of more than 30 hours and observed a strong increase in speed to 900km/s followed two hours later by a strong increase in the level of scintillation, interpreted as a strong increase in density. Both speed and density remained enhanced for a period of more than seven hours, to beyond the period of observation. Further analysis of these data demonstrates a view of magnetic-field rotation due to the passage of the CME, using advanced IPS techniques only available to a unique instrument such as LOFAR.
... The CCS in 2007 was observed in the solar wind that could not be attributed to a well-formed coronal hole; however, the existence of open field lines assured the high-speed outflow from the polar region for a prolonged time in 2007 February. The difference between two types of the fast solar wind from high and mid-heliolatitudes associated with coronal holes has also been discussed in Bisi et al. (2007) in terms of comparisons of Ulysses measurements and IPS data. ...
Article
Full-text available
We provide observational evidence for the existence of large-scale cylindrical (or conic-like) current sheets (CCSs) at high heliolatitudes. Long-lived CCSs were detected by Ulysses during its passages over the South Solar Pole in 1994 and 2007. The characteristic scale of these tornado-like structures is several times less than a typical width of coronal holes within which the CCSs are observed. CCS crossings are characterized by a dramatic decrease in the solar wind speed and plasma beta typical for predicted profiles of CCSs. Ulysses crossed the same CCS at different heliolatitudes at 2–3 au several times in 1994, as the CCS was declined from the rotation axis and corotated with the Sun. In 2007, a CCS was detected directly over the South Pole, and its structure was strongly highlighted by the interaction with comet McNaught. Restorations of solar coronal magnetic field lines reveal the occurrence of conic- like magnetic separators over the solar poles in both 1994 and 2007. Such separators exist only during solar minima. Interplanetary scintillation data analysis confirms the presence of long-lived low-speed regions surrounded by the typical polar high-speed solar wind in solar minima. Energetic particle flux enhancements up to several MeV/nuc are observed at edges of the CCSs. We built simple MHD models of a CCS to illustrate its key features. The CCSs may be formed as a result of nonaxiality of the solar rotation axis and magnetic axis, as predicted by the Fisk–Parker hybrid heliospheric magnetic field model in the modification of Burger and coworkers.
... Although these April observations correspond to a P-point location at middle-high latitudes, the LOS passes by the solar north pole during this month; the high latitude streams can contribute to the apparent solar wind speed. This result seems in general agreement with the magnitude and fluctuations of solar wind velocities (Smith (2011), and references therein) at high latitudes associated with solar cycle effects, and is also consistent with mid-latitude IPS fast solar wind reported by Bisi et al. (2007). ...
Article
Full-text available
Interplanetary scintillation (IPS) is used to probe solar wind speeds in the inner heliosphere by applying either of two generalized data-analysis techniques: model fitting to power spectra (MFPS) from a single station, or cross-correlation functions (CCF) produced by cross-correlating two simultaneous IPS time series from separate stations. The MEXican Array Radio Telescope (MEXART), observing at 140 MHz, is starting to use an MFPS technique. Here we report the first successful solar wind speed determinations with IPS observations by MEXART. Three stations of the Solar-Terrestrial Environment Laboratory (STEL), observing at 327 MHz, use a CCF, and an MFPS technique is also used at one of these sites. We here analyze data from MEXART and from one antenna of STEL to obtain solar wind speeds using an MFPS technique from a single station. The IPS observations were carried out with radio source 3C48 during Solar Cycle 24. The MFPS method we describe here is tested by comparing its obtained speeds with those from the STEL CCF technique. We find that the speeds from the two techniques generally agree within the estimated errors.
Article
Observations of interplanetary scintillation (IPS – the scintillation of compact radio sources due to density variations in the solar wind) enable the velocity of the solar wind to be determined, and its bulk density to be estimated, throughout the inner heliosphere. A series of observations using the Low Frequency Array (LOFAR - a radio telescope centred on the Netherlands with stations across Europe) were undertaken using this technique to observe the passage of an ultra-fast CME which launched from the Sun following the X-class flare of 10 September 2017. LOFAR observed the strong radio source 3C147 at an elongation of 82 degrees from the Sun over a period of more than 30 hours and observed a strong increase in speed to 900 km s⁻¹ followed two hours later by a strong increase in the level of scintillation, interpreted as a strong increase in density. Both speed and density remained enhanced for a period of more than seven hours, to beyond the period of observation. Further analysis of these data demonstrates a view of magnetic-field rotation due to the passage of the CME, using advanced IPS techniques only available to a unique instrument such as LOFAR.
Article
Observations of interplanetary scintillation (IPS) provide a set of data that are used in estimating the solar wind parameters with reasonably good accuracy. Various tomography techniques have been developed to deconvolve the line-of-sight integration effects ingrained in observations of IPS to improve the accuracy of solar wind reconstructions. Among those, the time-dependent tomography developed at the University of California, San Diego (UCSD) is well-known for its remarkable accuracy in reproducing the solar wind speed and density at Earth by iteratively fitting a kinematic solar wind model to observations of IPS and near-Earth spacecraft measurements. However, the kinematic model gradually breaks down as the distance from the Sun increases beyond the orbit of Earth. Therefore, it would be appropriate to use a more sophisticated model, such as a magnetohydrodynamics (MHD) model, to extend the kinematic solar wind reconstruction beyond the Earth's orbit and to the outer heliosphere. To testthe suitability of this approach, we use boundary conditions provided by the UCSD time-dependent tomography to propagate the solar wind outward in a MHD model and compare the simulation results with in situ measurements and also with the corresponding kinematic solution. Interestingly, we find notable differences in proton radial velocity and number density at Earth and various locations in the inner heliosphere between the MHD results and both the in situ data and thekinematic solution. For example at 1 AU, the MHD velocities are generally larger than the spacecraft data by up to 150 km s−1, and the amplitude of density fluctuations is also markedly larger in the MHD solution. We show that the MHD model can deliver more reasonable results at Earth with an ad hoc adjustment of the inner boundary values. However, we conclude that the MHD model using the inner boundary conditions derived from kinematic simulations has little chance to match IPS and textitin situ data as well as the kinematic model does unless it too is iteratively fit to the observational data and measurements.
Article
Full-text available
[1] We present results of observations of interplanetary scintillation (IPS) made using the telescopes of the MERLIN and EISCAT networks in which the beam separation approached 2000 km, much larger than in any previous IPS experiments. Significant correlation between the scintillation patterns was observed at time lags of up to 8 s and fast and slow streams of solar wind were very clearly resolved. One observation showed clear evidence of two discrete modes of fast solar wind, which we interpret as originating in the crown of the northern polar coronal hole and in an equatorward extension of the polar hole. We suggest that experiments of this type will provide a new and important source of information on the temporal and spatial variation of small-scale turbulence in the solar wind. The improved velocity resolution available from extremely long baseline measurements also provides new information on the development of the large-scale velocity structure of the solar wind in interplanetary space.
Article
The technique of interplanetary scintillation (IPS) is the observation of rapid fluctuations of the radio signal from an astronomical compact source as the signal passes through the ever-changing density of the solar wind. Cross-correlation of simultaneous observations of IPS from a single radio source, received at multiple sites of the European Incoherent SCATter (EISCAT) radio antenna network, is used to determine the velocity of the solar wind material passing over the lines of sight of the antennas. Calculated velocities reveal the slow solar wind to contain rapid velocity variations when viewed on a time-scale of several minutes. Solar TErrestrial RElations Observatory (STEREO) Heliospheric Imager (HI) observations of white-light intensity have been compared with EISCAT observations of IPS to identify common density structures that may relate to the rapid velocity variations in the slow solar wind. We have surveyed a one-year period, starting in April 2007, of the EISCAT IPS observing campaigns beginning shortly after the commencement of full science operations of the STEREO mission in a bid to identify common density structures in both EISCAT and STEREO HI datasets. We provide a detailed investigation and presentation of joint IPS/HI observations from two specific intervals on 23 April 2007 and 19 May 2007 for which the IPS P-Point (point of closest approach of the line of sight to the Sun) was between 72 and 87 solar radii out from the Sun’s centre. During the 23 April interval, a meso-scale (of the order of 105 km or larger) transient structure was observed by HI-1A to pass over the IPS ray path near the P-Point; the observations of IPS showed a micro-scale structure (of the order of 102 km) within the meso-scale transient. Observations of IPS from the second interval, on 19 May, revealed similar micro-scale velocity changes, however, no transient structures were detected by the HIs during that period. We also pose some fundamental thoughts on the slow solar wind structure itself.
Article
The technique of interplanetary scintillation (IPS) can be used to probe interplanetary space between the Sun and Earth most-commonly in terms of speed and also by using the scintillation-level (g-level) as a proxy for density. We combine the large spatial-scale 3D tomographic techniques previously only applied to IPS data from the Solar Terrestrial Environment Laboratory (STELab) array, Nagoya University in Japan, and the previously operational Cambridge IPS system in England, with the finer-scale capabilities of the longer baselines between the systems of the Multi-Element Radio-Linked Interferometer Network (MERLIN) in the UK, and the European Incoherent SCATter (EISCAT) radar and the EISCAT Svalbard Radar (ESR) in northern Scandinavia. Using the UCSD D reconstruction technique, we present results of detailed measurements of speed in the solar wind and also those of solar wind flow-directions, constrained by the large-scale density tomography through the use of a kinematic model, as well as applying this tomographic technique for the first time to the MERLIN, EISCAT, and ESR IPS solar wind speed observations in terms of velocity.
Article
Full-text available
Improvements to two of the radio telescopes of European Incoherent SCATter radar (EISCAT) allow measurements of Interplanetary Scintillation (IPS) at 1.4 GHz, and this has prompted two major developments in studies of IPS. Simultaneous observations between EISCAT and MERLIN allow baselines of up to 2000 km, significantly improving velocity resolution and making possible much greater accuracy in determining the direction of flow than previously, as well as demonstrating that density variations in the slow solar wind remain partially correlated for at least 8 s. Initial results suggest two fast solar wind modes, and there is evidence for super-radial (meridional) expansion of the fast solar wind. Trials have been conducted using different observing frequencies at different sites, including the use of the EISCAT Svalbard Radar for IPS for the first time. Observations at 500, 928 and 1420 MHz with baselines of up to 1200 km have been carried out. The results are found to be consistent between single- and dual-frequency correlations, allowing the range of observations possible with the EISCAT system to be expanded.
Article
Full-text available
[1] We present results of observations of interplanetary scintillation (IPS) made using the telescopes of the MERLIN and EISCAT networks in which the beam separation approached 2000 km, much larger than in any previous IPS experiments. Significant correlation between the scintillation patterns was observed at time lags of up to 8 s and fast and slow streams of solar wind were very clearly resolved. One observation showed clear evidence of two discrete modes of fast solar wind, which we interpret as originating in the crown of the northern polar coronal hole and in an equatorward extension of the polar hole. We suggest that experiments of this type will provide a new and important source of information on the temporal and spatial variation of small-scale turbulence in the solar wind. The improved velocity resolution available from extremely long baseline measurements also provides new information on the development of the large-scale velocity structure of the solar wind in interplanetary space.
Article
Full-text available
We investigate the extension and evolution of the solar corona into interplanetary space by comparing 1995 Ulysses radio occultation measurements of path-integrated electron density and density fluctuations measured between 21 and 32 Ro, with simultaneous white-light measurements made by the HAO Mauna Loa K-coronameter below 2.5 Ro. The surprising picture of the extended corona to emerge from this comparison is one in which stalks of streamers, occupying a small fraction of volume in interplanetary space, are superimposed on a background corona distinguished by a plethora of ray like structures, often referred to as plumes in polar coronal holes. The radial preservation of the boundary between polar coronal holes and the base of streamers implies that the solar wind from polar coronal holes expands radially rather than undergoing any significant divergence as previously thought. Combining this picture of the extended corona with in situ velocity measurements made by Ulysses throughout its two polar passages, we conclude that the raylike structures, except for the stalks of streamers, seem to be the source of the fast wind. The existence of the fast wind at low latitudes can be attributed to these raylike structures, rather than the expansion of the boundaries of polar coronal holes to low latitudes.
Article
Full-text available
The ESA/NASA spacecraft Ulysses provides the first in-situ observation of the solar wind at high solar latitudes. Data obtained with the Solar Wind Ion Composition Spectrometer (SWICS) during the first orbit of Ulysses around the Sun can be used to obtain a global view of the solar wind pattern within the heliosphere during the declining/minimum phase of the solar activity cycle. This provides information on the solar corona, specifically on any latitudinal variation of coronal parameters as far as they play a role in the solar wind acceleration process. The measurements show: The 3-dim. distribution of the solar wind appears as a pattern of basically two states. The fast solar wind with speeds of about 750 km/s emanating from the polar coronal holes dominates the heliosphere. The slow solar wind from the streamer belts is restricted to a region of about +/−20° around the heliographic equator. The boundary between the two states of the solar wind is sharply defined. The coronal hole streams are remarkably stable and homogeneous on a larger scale. The two independent coronal holes are similar in terms of average speed and spatial extent. In both coronal holes a gradual increase (≤10%) of speed with solar latitude is observed.
Article
Until the ULYSSES spacecraft reached high latitude, the only means for measuring the solar wind velocity in the polar regions was from radio scattering observations (IPS), and these remain the only way to measure the velocity near the sun. However, IPS, like many remote sensing observations, is a “line-of-sight” integrated measurement. This integration is particularly troublesome when the line-of-sight passes through a fast stream but that stream does not occupy the entire scattering region. Observations from the HELIOS spacecraft have shown that the solar wind has a bimodal character which becomes more pronounced near the sun. Recent observations from ULYSSES have confirmed that this structure is clear at high latitudes even at relatively large solar distances. We have developed a method of separating the fast and slow contributions to an IPS observation which takes advantage of this bimodal structure. In this paper I will describe the technique and its application to IPS observations made using the receiving antennas of the EISCAT incoherent backscatter radar observatory in northern Scandinavia.
Article
We study the properties of micro-turbulence in the solar wind using interplanetary scintillation (IPS) observations made with the single dish antenna operating at 2.3 GHz and 8.5 GHz. Our IPS observations were made during September-October in 1992 and 1993 and covered a radial distance range of 5-76 solar radii (Rs). We apply the spectral-fitting method to obtain properties of solar wind turbulence, such as axial ratio of anisotropy, inner scale of dissipation length, power-law spectral index, and solar wind velocity. We also examine the velocity dependence of the turbulence near the sun, and find the following results. (1) Both low-speed and high-speed solar-wind flows show acceleration at the distance range of 10-30 Rs, which is consistent with previous works. (2) The radial dependence of anisotropy and spectral index show no significant difference between the low-speed and high-speed solar wind. (3) Only the inner scale length shows a dependence on flow speed: the inner scale in the high-speed wind is 1.4 ± 0.3 times larger than that in the low-speed wind.
Article
The EISCAT 930 MHz antennas have for many years been used for interplanetary scintillation measurements of solar wind parameters. To augment the EISCAT IPS capabilities, the Swedish and Finnish antennas were recently equipped with 1400 MHz dual polarisation receiver systems, which in addition to the total signal power can also measure phase and amplitude scintillations simultaneously. This is potentially very important, as amplitude scintillation measurements provide better information on solar wind velocity while phase scintillation measurements are a particularly powerful method of studying the solar wind turbulent-scale microstructure. The first 1400 MHz IPS campaign is planned for May 2002.
Article
We have studied the radial dependence of the velocity of high-latitude fast solar wind in the heliocentric distance range of 0.13–0.9 AU. For this study a new tomographic analysis method which can evaluate uncertainties was developed to obtain velocity distribution maps on two reference spheres at 0.13 and 0.3 AU using interplanetary scintillation (IPS) observations. First of all, it is tested that this tomographic method has enough sensitivity and reliability to investigate the radial dependence of the wind velocity. The analysis was made for the IPS observations during 3 years, from 1995 to 1997, when solar activity was minimum. From this analysis, average velocities of 770–780 km s−1 were obtained at distances of 0.13–0.3 AU, which were 19 ± 17 km s−1 lower than those at 0.3–0.9 AU. The results from this work, taken together with measurements of SOHO/LASCO, EISCAT and MERLIN [Breen et al., 2002], Helios [Schwenn et al., 1978], and Ulysses [McComas et al., 2000], indicate that the fast wind is accelerated almost to its final flow velocity within 20 Rs and a small but not negligible acceleration exists beyond 30 Rs which tends to become smaller at farther heliocentric distances.
Article
This study examines solar wind plasma and magnetic field observations from Ulysses' first full polar orbit in order to characterize the high-latitude solar wind under conditions of decreasing and low solar activity. By comparing observations taken over nearly all heliolatitudes and two different intervals covering the same radial distances, we are able to separate the radial and latitudinal variations in the solar wind. We find that once the radial gradients are removed, none of the high-latitude solar wind parameters show much latitudinal variation, indicating that the solar wind emanating from the polar coronal holes is extremely uniform. In addition, by examining nearly 6 years of data starting in the declining phase of the last solar cycle and extending through the most recent solar minimum, we are able to address hemispheric asymmetries in the observations. We find that these asymmetries are most likely driven by differences in the solar wind source over the solar cycle and indicate that more energy goes into the polar solar wind during the declining phase of the solar cycle than around minimum. Because the mass flux is larger in the declining phase while the speeds are very similar, we conclude that this energy is introduced at an altitude below the solar wind acceleration critical point. Finally, we provide details of the statistics of over 20 solar wind parameters so that upcoming observations from Ulysses' second polar orbit, during much more active times on the Sun, can be readily compared to the quieter first orbit results.
Article
Observations of the coherence bandwidth of interplanetary scintillations (IPS) made at the Nançay Radio Observatory between 1982 and 1992 have been used to estimate the variance of electron density deltaN2e at solar distances of 5 to 15 Rs. These IPS observations are sensitive to scales of the order of 10 km; i.e., they measure the ``microscale variance,'' which is a small fraction of the total density variance. Measurements in the equatorial region are invariant over the solar activity cycle and show the same behavior with distance as the mean density, i.e., deltaN2e~N2e. They do not show a statistically significant change in the ``transonic region'' as reported by other observers (Lotova et al., 1985). Measurements over the polar region show a clear variation with the solar cycle. At solar maximum, deltaN2e is invariant with latitude, as is Ne (Withbroe, 1988). However, the apparent deltaN2e measured by IPS is about 10 times lower in the polar streams which form during the declining and minimum phases of solar activity. Compensation for the line of sight integration inherent in the IPS observations indicates that the local value of deltaN2e is actually about 15 times lower in the polar streams. We cannot estimate the latitude variation of the ``development'' of the compressive turbulence deltaNe/Ne between 5 and 15 Rs because measurements of Ne in the polar hole are lacking.
Article
Statistical models used in the analysis of drifting random patterns are discussed. A new method of estimating the drift velocity with minimal a priori assumptions is presented. It is shown that all the models which allow for pattern rearrangement, including the new method, can be highly sensitive to systematic errors. These techniques are applied to recent observations of interplanetary scintillation in the ecliptic plane. A comparison is made with mean proton velocities measured by satellite, and a clear correspondence between high velocity features is found. Measurements of the scintillation of smalldiameter radio sources caused by the solar wind can be used to derive information about the source structure and to measure some of the parameters of the solar wind [Dennison and Hewish, 1967; Cohen et al., 1967; Dennison, 1969; Hewish and Symonds, 1969]. The purpose of this paper is to discuss the methods of analysis of three-station interplanetary scintillation (IPS) that can be applied to the latter problem. The results of recent 74-MHz measurements made with the University of California. solarwind observatory at San Diego are given. The measurable process in IPS is the intensity of the diffraction pattern caused by random variations in the electron density of the solar wind. The analysis problem is to deduce statistics of the variations in electron density from observations of the time variation of the pattern at several spaced stations. The approach used has been to fit a statistical model to the measured pattern and then to relate this model to an appropriate model of the solar wind. Model parameters normally include a 'spatial scale,' a velocity, and a 'rearrangement time' (or random velocity component). One problem with this approach is that the solar wind may not have a well-defined scale or velocity. A spatial scale implies an 'effective width' of the spatial power spectrum, but if this spectrum has a power-law shape with wave number, then such a scale cannot be defined. The diffraction pattern itself must have a measCopyright ¸ 1972 by the American Geophysical Union. urable 'scale,' since the low spatial frequencies are attenuated by the effect of propagation, the so-called Fresnel filter; and the high frequencies are limited either by some plasma cutoff frequency or the source diameter. The pattern velocity can be defined as the velocity of an inertial frame in which the pattern is stationary. In this frame, there is no correlation between temporal and spatial variations, so that the covariance functions must be symmetric. Such a velocity cannot exist, for example, if the scattering medium consists of two distinct regions with different velocities and densities. However, if there are a large number of regions and the distribution of velocities is symmetrical, a velocity can be defined.