349
1 INTRODUCTION
The utilisation of PNT services is increasingly taking
place in current and planned applications where high-
precision and integrity of services are required. It
reflects especially on critical applications in maritime
navigation and transportation branches in general.
Developing applications are requiring demanding
tasks of sub-meter accuracy such as maritime
operations, traffic management, search and rescue
operations, marine engineering, as well as offshore
and port systems operations. Furthermore, satellite
positioning represents the primary navigational data
source onboard vessels. The timing data accuracy is
essential for the management of sailing passages,
ports and approaches and navigational lights
synchronisation (Thomas et al., 2011; Brčić, 2012).
Performance degradation of GNSS services can affect
certain system/application to a greater or lesser extent,
depending on the level of required availability,
integrity and accuracy (Thomas et al., 2011).
Satellite positioning performance is susceptible to
errors caused by a number of individual causes,
Reconstruction of Geomagnetic Event as Observed in Northern
Adriatic Region and Its Correlation with GPS Single
-frequency
Positioning Deviations
D. Br
čić, J. Ćelić & S. Valčić
University of Rijeka, Rijeka, Croatia
ABSTRACT: Space weather effects are generally recognized as causes of degradation of satellite positioning,
navigation and timing (PNT) services. We analyze GPS position estimation error during a geomagnetic storm,
focusing on manifestations of geomagnetic processes. The position estimation error was analyzed in terms of
GPS coordinates’ deviations (latitude, longitude and height) from their reference values. The storm’s impact
was studied in the Northern Adriatic region where GPS observables from two Global Navigation Satellite
System (GNSS) reference stations were analysed. Geomagnetic indices were elaborated, comprising readings
from interplanetary, magnetospheric and geomagnetic observatories. Total Electron Content (TEC) on both
stations was computed using dual frequency GPS pseudorange observables. The experiment was to reconstruct
the movement of geomagnetic disturbances entering the geospace, reaching the earth’s surface. The aim was to
correlate possible space weather manifestation on satellite positioning performance in terms of positioning
error. Regularities in changes in positioning deviations were identified with relation to influential indices. The
research offered a possibility of experimental positioning deviations assessment as well as forecasting.
Evaluation of generated rudimentary Classification and Regression Trees (CART) models showed that the risk
of satellite positioning errors could be assessed and predicted considering absolutes, as well as changes in
values of geomagnetic indices. During the research process, several activities emerged as preferable
continuation of the work, with the aim of further development of predictive models and the complement of
space weather scenarios and their consequences on navigational systems. Along with summarized results, they
are outlined in the conclusion section.
http://www.transnav.eu
the
International Journal
on Marine Navigation
and Safety of Sea Transportation
Volume 14
Number 2
J
une 2020
DOI:
10.12716/1001.14.02.11
350
varying from spatial distribution of satellites to user’s
micro-environment. Signal propagation error
represents the main cause of single-frequency
positioning degradation (Kintner and Ledvina, 2005;
Parkinson and Spilker, Jr., 1996) with ionospheric
delay as a prevalent positioning error cause
(Klobuchar, 1983, 1987). Ionospheric delay is
proportional to the amount of TEC on the path
between the satellite and receiver antenna (Klobuchar,
1988). TEC is expressed in TEC Unit (TECU), one
being 10
16
e / m
2
, equivalent to 1.624 meters of
measured pseudorange error at GPS frequency L1
(Parkinson and Spilker, Jr., 1996).
TEC behavior varies depending on different
occurrences and at different timescales (Mendillo,
2006); diurnal and seasonal variations, solar cycle and
storm time behavior. The motivation of the conducted
research was an attempt to create an image of near-
earth space environment during disturbed space
weather, before geomagnetic disturbances reach
GNSS orbit heights, with eventual positioning
deviations as final manifestation. A geomagnetic
event was reconstructed to determine the relation
between the geomagnetic storm effects and GPS
positioning error. Data describing geomagnetic
indices were analysed together with positioning
observables on two International GNSS Service (IGS)
stations in the Northern Adriatic area. Regularities of
positioning patterns were detected as a response to
the storm and variations in the geomagnetic field.
Subsequently, the ionospheric analysis was made
with local TEC behavior as a linkage between
magnetic variations and the final positioning
deviation. Event reconstruction was presented with
the objective to confirm geomagnetic impact on
satellite positioning performance. After analysed data
interpretation, modeling of latitude, longitude and
height deviations was introduced and evaluated. In
the concluding chapter, research results and
inferences were outlined along with planned activities
which emerged from the research.
2 BACKGROUND
Satellite navigation technology presents an inevitable
path towards improvements of transportation as a
master link between sustainability pillars (Brčić,
2012). PNT services are used in all transportation
branches, whether in traffic control management,
monitoring or navigation. The European Union Single
European Transport Area sub-projects are designed to
ease the citizens and cargo movements and to
enhance the European transport sustainability;
however, the vision is global. Current and planned
strategies imply an optimal infrastructure and reliable
sources of dynamic information regarding PNT data
provision (Thomas, 2011).
User Equivalent Range Error (UERE) encompasses
influential quantities which are affecting time
measurement of the satellite signal propagation,
leading to erroneous pseudorange computation
(Parkinson and Spilker, Jr., 1996; Subirana et al., 2013,
Kaplan and Hegarty, 2006):
( )
uD
rct t t
ρ δδ
=+⋅ +
(1)
where:
... pseudorange between satellite and user
antennae,
r
... geometric (true) range between
satellite and user antennae,
u
t
... user clock error,
t
δ
... satellite clock offset,
D
δt
... the total ranging
error generated by other influential factors.
The range measurement timing relations are
shown in Figure 1, as interpreted from (Kaplan and
Hegarty, 2006).
The ionospheric delay can be expressed as:
40.3/ ^ 2f TEC
ρ
∆=
(2)
where
ρ
equivalent ionospheric delay of
determined pseudorange,
f
system’s operating
frequency,
TEC
Total Electron Content along an
equivalent column between satellite and user’s
antenna with a cross section of 1 m
2
.
Figure 1. Range measurement timing relations
Influences on GPS positioning error during
disturbed space weather have already been
investigated in the area (Filjar et al., 2013), together
with the development of GNSS positioning
performance forecasting (Filić and Filjar 2018).
Ionospheric disturbances are the consequence of
changes, variations and occurrences resulting from
conditions in the solar-terrestrial environment due to
the manifestation of the solar activity (Parkinson and
Spilker, Jr., 1996; Goodman, 2005). According to
National Oceanic and Atmospheric Association’s
Space Weather Prediction Center (NOAA SWPC),
geomagnetic storms, solar radiation storms and radio
blackouts are described by numbered levels and with
the corresponding indicators according to severity.
Geomagnetic activity is described by indices
presented in Table 1 (Perrone and De Franceschi,
1998; Zolesi and Cander, 2014). Solar eruptive events
create magnetic disturbances Coronal Mass
Ejections (CME) which travel through interplanetary
space and interact with thermosphere, ionosphere and
magnetosphere (Lockwood et al., 1999). The planetary
Kp and Ap indices express the horizontal component
of the geomagnetic stability. There appears a negative
correlation between the Kp index and the storm-time
TEC during summer months (Mendillo, 2006; Ross,
1960). The southward orientation of earth-directed
interplanetary magnetic field (IMF) expressed with
the B
Z index causes the entering of power inputs in
351
the magnetosphere, producing geomagnetic
disturbances (McMorrow, 2011).
Table 1. Geomagnetic indices
_______________________________________________
Index Description
_______________________________________________
K/Kp Three-hour pseudo-logarithmic index representing
disturbances in horizontal component of the
geomagnetic field in relation to quite space weather
conditions.
Dst One-hour indicator of magnetospheric activity, as a
measure of the ring current in the magnetosphere.
It can be interpreted as longitudinal average of
horizontal geomagnetic disturbance.
AE One-hour indicator of auroral (electrojet) activity,
further divided in amplitude upper (AU) and
amplitude lower (AL) indices with the relation
AE=AU-AL A/Ap Linear equivalent of the K/Kp
index, representing average of geomagnetic field
variations
aa Global geomagnetic activity index derived from
two mutually antipodal magnetic observatories.
BT Geomagnetic field intensity indicator, which can be
further decomposed into northing, easting and
vertical component.
_______________________________________________
The Bz index is one of the first qualitative storm
indicators, as its southward pointing triggers
processes in the geospace, including an opening of the
magnetosphere (Lockwood et al., 1999). The
magnetospheric opening causes increased ring
current (Dst), allowing the energy and particles to
enter through auroral ovals (AE/AU/AL) and spread
over the globe (Mendillo, 2006; Booker, 1954). These
processes lead to disturbances in the ionospheric F2
region and on distribution and behavior of the total
electron content (Davies, 1965).
3 METHODOLOGY
The first group of analysed parameters refers to
geomagnetic indices, while the second represents
three-dimensional satellite positioning deviations
obtained with GPS. Readings of magnetic field
changes in the magnetosphere were retrieved from
the Geostationary Operational Environmental
Satellite (GOES) 15 measurements archive (Singer et
al., 1996). Apart from geostationary observations,
readings of magnetospheric field components were
retrieved from Advanced Composition Explorer
(ACE) spacecraft observables, measured at Libration 1
point.
Earth observations were analysed near to the
Northern Adriatic region, using nearest of
International Real-Time Magnetic Observatory
Network (INTERMAGNET) observatories (Chambon
la Foret, France). Global geomagnetic indices Dst, AE,
AU, AL and Kp, were retrieved from the World Data
Center for Geomagnetism (WDC) and Space Physics
Interactive Data Resource (SPIDR) databases.
Collection, processing, determination and creation of
GPS positioning solutions were enabled using Reader
Independent Exchange Format (RINEX) databases,
available at International GNSS Service and US NGS
CORS servers (Gurtner and Estey, 2009). Data from
two stations were analysed: Bolzano, Italy and Graz,
Austria (Table 2). It is assumed that the consistency
and quality of IGS data is ensured through (Kouba,
2009), focusing on the removal of non-dispersive
pseudorange error components, such as multipath (<
0.3 m), cycle slips (< 1 per 1000 observations),
northing, easting and height eccentricities to the
antenna reference point (≤1 mm), and other
requirements needed for the provision of high-quality
and high-integrity GNSS products.
Table 2. General information of Bolzano and Graz IGS
stations
_______________________________________________
ID City Location Longitude Latitude Height
(E) (N) (m)
_______________________________________________
bzrg Bolzano Italy 11.3368 46.4990 328.8
graz Graz Austria 15.4935 47.0671 538.3
_______________________________________________
GPS positioning files (RINEX.pos) were calculated
and archived using RTKLIB open source program
package, employing observation and navigation files.
The ionospheric correction was settled by employing
standard ionospheric model, coefficients of which
were taken from diurnal navigational messages.
Other standard methods and algorithms were settled
as follows: single positioning mode, L1 frequency
(single) positioning solution, 15° elevation mask,
broadcast ephemeris, continuous ambiguity
resolution and tropospheric model (Saastamonien).
Ionospheric TEC was calculated by using dual-
frequency (L1 and L2) GPS measurements from IGS
stations. Measured phase differences (
TEC
φ
) were
smoothed with code differences (
P
TEC
), after which
they were leveled again using code differences to
correct TEC values (Dyrud et al. 2008):
21
1
0.104
P
PP
TEC
mTECU
=
(3)
12
1
0.104
TEC
mTECU
φ
φφ
=
(4)
After Differential Code Biases (DCB) correction
(Noll, 2010), it is assumed that all frequency-
dependent quantities except ionospheric delay were
eliminated (Subirana et al., 2013). Standard deviations
of computed values were presented as well. The total
electron content was derived using GPS-TEC software
from IGS stations’ RINEX files, by using described
mathematical relations.
Time series were analyzed from all parameters,
along with statistical summaries and distribution of
collected data. The event phases were further divided,
isolating specific influences and different behavior of
indices (e.g. depletion and enhancement of TEC).
Correlation between variables was made using
Pearson’s correlation coefficients for the two most
pronounced observed events.
Recursive binary splitting (partitioning) algorithm
was used to develop a decision tree model, which the
main objective is to minimise the Residual Sum of
Squares (RSS) given as (Hastie et al., 2009):
352
( )
2
1
ˆ
m
M
im
m iR
RSS y yR
=
=
∑∑
(5)
where:
M
... partitions of partitioned feature space,
i
y
... response of a particular testing observation, and
ˆ
m
yR
... mean response of the training observation
within partition
m
.
Development of such models is based on the
concept of splitting the observed data set into subsets
and on the values of predictors in a series of iterations
(Filić and Filjar, 2018).
For the sake of simplicity, the procedure was
limited to the determined depth of a tree. The general
model is given as follows (Hastie et al., 2009, James et
al., 2013):
(
) (
)
( )
1
|;
M
m
m
f yw
φ
=
= =
xx x
m
v
(6)
where:
m
w
the mean response for the
particular region
m
R
, and
m
v
represents how each
variable is split at a particular threshold value.
During splitting, Gini index as the main criteria
was used to assess the function (Hastie et al., 2009,
James et al., 2013):
( )
2
1
1 (|)
=
=
k
j
it p jt
(7)
where:
k
the number of possible output
categories, while the category
j
has a probability of
occurrence
( |)pjt
.
After several algorithm repetitions, the final
selection of indices was made, also considering the
correlation results. Regression trees as predictive
models were generated for each positioning
component as target variables. The final output of
models was the estimation of probable values of
latitude, longitude and height deviations,
respectively. Proposed models were evaluated
employing observed and predicted values.
Partitioning of used data-sets was divided as 70% for
training, 15% for validation and 15% for evaluation,
respectively.
4 THE EVENT
In this section, the geomagnetic event was described,
as occurred through June 21 27, 2015 following
official event reports, using available data and
followed with own interpretation.
4.1 Geomagnetic storm development
According to the United States Geological Survey
(USGS) National Geomagnetism report, the storm
occurred because of solar wind fast stream and CMEs,
arriving on June 21
st
at 16:45 UT (day of year (DOY)
172), June 22
nd
at 05:45 UT (DOY 173), and June 22
nd
at
18:30 UT, compressing the magnetosphere and
generating electric currents and geomagnetic field
perturbations.
Figure 2. IMF Bz component magnitude 4-minue level data
(top) and Kp index (bottom) 3-hours data through the storm
period, DOY 172-177
In a specific moment, IMF Bz component turned
southward (Figure 2). Geomagnetic activity arose
firstly at high latitudes, as shown in Figure 3. The
greatest Dst depression occurred on June 23
rd
at 04:30
UT and recovering until another CME reached the
earth (Figure 3). The sub-events formed geomagnetic
storm period of seven days. Statistical parameters
were calculated based on the time series for the
specific indices, as presented in Table 3.
Figure 3. Auroral activity and Dst values, DOY 171 – 181
According to (Singer et al., 1996), nominal
(undisturbed) total geomagnetic field intensity (BT)
measured on satellite vary from -200 nT to +200 nT.
The BT maximum value and range exceeds the limits
several times.
353
Table 3. Statistical description of geomagnetic indices
during DOY 172-177
_______________________________________________
Index Min Mean SD Max Range
_______________________________________________
GOES BT 33.93 139.5135 72.2918 1,289.21 1,255.28
(nT)
Dst (nT ) -204.00 -46.34 41.20028 36.00 -240
AE (nT) 31.00 280.30 296.5678 1,636.00 1,605
AU (nT) 6.0 115.7 123.0969 841.0 835
AL (nT) -1,101.0 -164.6 190.7277 -13.0 1,088
Kp 0.300 3.273 8.300 1.9665 8.0
CLF BT 47,741.3 47,824.5 25.6576 47,976.3 234.96
(nT)
_______________________________________________
Pronounced values indicate geomagnetic storm
and travelling disturbances and high variations. The
Kp increased two times; the maximum of 8.3 took
place just before the commencement of day 174 (June
23
rd
), associated with June 21
st
and June 22
nd
CMEs.
The second increase occurred on day 176 (June 25
th
)
after another solar eruption and CME arrival,
respectively.
4.2 Storm-time GPS positioning deviations
Time series of deviations of positioning components
were calculated for the period June 22
nd
(DOY 173)
June 26
th
UTC (DOY 177). Horizontal deviation plots
are shown in Figure 4.
Figure 4. Horizontal positioning error as calculated for
stations Bolzano (top) and Graz (bottom).
Statistical description of GPS positioning error
through the observed period is shown in Table 4, with
arc units converted to meters.
Table 4. Bolzano (B) and Graz (G) horizontal positioning
error during the storm period
_______________________________________________
Station Lat error (m) Long error (m) Height error
(m)
Mean SD Mean SD Mean SD
_______________________________________________
G 4.191 1.017 -1.350 0.546 1.572 1.849
B 3.524 1.028 -0.126 0.6 0.098 1.875
_______________________________________________
The distribution of observed height values for both
stations is presented in Figure 5.
Although there is a constant offset relative to
reference values, latitude and longitude are
experiencing smaller deviations when compared to
height, tending northward (latitude) and eastward
(longitude), respectively. Standard deviations are
approximately the same at both locations and in all
axes; however, vertical component shows the most
prominent scattering.
Figure 5. Height histograms for Graz (top) and Bolzano
(bottom)
Statistical description of daily positioning
deviations is graphically presented in Figure 6,
showing minimal and maximal values and range,
mean values and standard deviation, respectively.
354
Figure 6. Bolzano (top) and Graz (bottom) daily horizontal
positioning error statistics
4.3 Interpretation of results
The TEC patterns during the storm period with
computed mean values and standard deviations are
shown in Figure 7.
Figure 7. TEC daily patterns mean values (top) and
standard deviations (bottom) as computed for locations
Bolzano and Graz during the storm period
In Table 5, statistical parameters of the computed
storm-time period TEC at locations Bolzano and Graz
are presented.
Table 5. Statistical description of the total electron content
_______________________________________________
Station Minimum Mean Standard Maximum Range
value value deviation value
_______________________________________________
Bolzano 2.09 10.36 5.59 28.55 26.46
Graz 2.03 9.78 5.51 26.57 24.54
_______________________________________________
On Figure 8, readings from GOES magnetometers
1 and 2 are presented, together with geomagnetic
field intensity as measured on Chambon-La-Foret
observatory. On Figure 9, combined plots of
positioning error deviations for Bolzano and Graz
locations are shown, together with associated TEC
computed on both locations. The illustrations are
followed with interpreted results.
Figure 8. Geomagnetic field intensity as measured with
Chambon-La-Foret (top), GOES 1 (middle) and GOES 2
(bottom) magnetometer
Abrupt change (depression) of TEC (DOY 173)
preceded the increase of geomagnetic field intensity.
After geomagnetic peak values and during TEC
decline (end of DOY 173 / beginning of DOY 174),
positioning deviations were most pronounced. At the
same time, the greatest Dst drop and AE increase
occurred. During first hours of day 174, a measurable
TEC peak occurred.
TEC commotions accompanied the second
positioning deviations increase on DOY 174. During
DOY 175, all the observed values were returning in
regular patterns. The increase of Dst indicated
355
possible commencement of new geomagnetic
disturbance.
The DOY 176 is characterized by Dst decline, an
increase of auroral activity and stirrings of all
(magnetospheric and ground) magnetic field
components. Total geomagnetic field flux measured at
GOES magnetometer 1 reached the greatest value
during this period. Second GOES magnetometer
shows oscillations of magnetic components around
the mean values, like ground readings. The drop of Bz
component as measured on the first GOES
magnetometer is here pronounced. Positional
deviations started to increase, reaching maximum
values before midday. The following hours were
characterised with TEC increase, reaching its greatest
observed values. The magnetospheric total field
strength was the highest observed.
The disturbance ceased the following day, which
can be interpreted from regular patterns of
geomagnetic components, as well as TEC lowering. In
line with described patterns, positioning deviations
are also decreasing, returning to normal values.
Figure 9. Graz (top) and Bolzano (bottom) latitude (red),
longitude (blue) and height (green) positioning deviation
patterns, with total electron content (black), computed at
each location
5 DISCUSSION
Two distinctive incidents (events) can be extracted:
abrupt depression and subsequent rise on DOY 173
and pronounced increase on DOY 176.
For these two periods geomagnetic, positioning
and TEC parameters were used to find mutual
correlations. The results are shown in Figure 10 and in
Figure 11. During the first event, the total magnetic
field intensity measured on all magnetometers
(BTCLF, BTG1 and BTG2) correlated positively with
latitude and height positioning errors (bzrg_fi, bzrg_h,
graz_fi, graz_h), while negative correlation were found
with longitude errors (bzrg_l, graz_l). The TEC
(bzrgtec, graztec) showed negative correlations with
components of GOES-measured geomagnetic activity,
and positive correlation with ground-based magnetic
readings. During the second event, BTG1 showed a
positive correlation with positioning components,
opposite to Chambon-La-Foret (ground) readings,
same being applicable for TEC.
Positioning deviations were most expressed
during pronounced changes in geomagnetic field
intensity components, measured with GOES
magnetometer 1 (BxG1, ByG1, BzG1, BTG1), GOES
magnetometer 2 (BxG2, ByG2, BzG2, BTG2) and
Chambon-La-Foret (BxCLF, ByCLF, BzCLF, BTCLF).
The TEC behavior showed correlation with
geomagnetic activity and the affectation of the
geomagnetic impact on both stations similarly.
Figure 10. Correlation matrix between variables during the
first event. The dots size indicates positive (white) or
negative (black) correlation.
Figure 11. Correlation matrix between variables during the
second event. The dots size indicates positive (white) or
negative (black) correlation.
The time of commencement and duration of the
main phase of the storm (DOY 173.5 174.5) was
extracted for modeling of positioning deviations of
356
components on station Graz: latitude (devFIG),
longitude (devLG) and height (devHG). During the
height model-building process, main components
were defined as AU, BTCLF, BxCLF, BzCLF, BzG1,
TEC, dTEC, and Dst. A similar process was made for
latitude and longitude models, respectively.
Governing indices were found as follows:
Latitude model: AE, AL, ByCLF, ByG1, ByG2,
dBzCLF, TEC, Dst,
Longitude model: AE, AU, BTG1, BTG2, ByCLF,
ByG1, BzCLF, TEC, Dst.
For the latitude target variable, the root node was
a northing magnetic component as measured on
second GOES magnetometer (ByG2), while the total
magnetic field intensity (BTG1) was the root node for
the longitude model.
Figure 12. Latitude (left), longitude (centre) and height
(right) deviations model evaluation
Decision tree models can develop certain
drawbacks in the form of extreme sensitivity and ease
of over-fitting. However, they represent a simple
decision-making tool. The success of models will
depend on the appropriate selection of possibly
influential quantities. As for the proposed models,
obtained results were compared with observed values
of positioning deviations. Evaluation of models is
shown in Figure 12. The plots are showing observed
versus predicted values, local regression scatterplot
smoothing, regression lines and confidence intervals.
Considering adjusted R-squared, the average
overall score of the efficiency of models is 71.4%
(Table 6).
The height model showed the best performance
(78%). In general, vertical satellite positioning
component is most susceptible to external influences.
The latitude model evaluation results showed 69%,
and longitude model showed 67% of performance
efficiency, respectively.
Table 6. Model evaluation output: observed against
predicted values regression analysis
_______________________________________________
Height deviation model
_______________________________________________
Coefficients Estimated Standard Pr(>|t|)
coefficient Error
_______________________________________________
Intercept 0.19438 0.06719 0.00421
Predicted deviation 0.96744 0.03458 < 2e-16
R-squared Multiple: 0.7845 Adjusted: 0.7835
F-statistic: 782.5 on 1 and 215 DF p-value: < 2.2e-16
_______________________________________________
Latitude deviation model
_______________________________________________
Coefficients Estimated Standard Pr(>|t|)
coefficient Error
Intercept -0.1192 0.3568 0.739
Predicted deviation 0.9921 0.04552 < 2e-16
R-squared Multiple: 0.6884 Adjusted:0.687
F-statistic: 475 on 1 and 215 DF p-value: < 2.2e-16
_______________________________________________
Longitude deviation model
_______________________________________________
Coefficients Estimated Standard Pr(>|t|)
coefficient Error
Intercept -0.1277 0.28384 0.653
Predicted deviation 0.9274 0.04413 < 2e-16
R-squared Multiple:0.6726 Adjusted:0.6711
F-statistic: 441.7 on 1 and 215 DF p-value: < 2.2e-16
_______________________________________________
6 CONCLUSIONS
The amount and intensity of space weather effects on
PNT services depend on an individual event. A
geomagnetic event which took place in June 2015 was
studied in this paper. The initial aim was the
reconstruction of the flow of geomagnetic occurrences
from magnetospheric heights up to the surface of the
earth, and to correlate geomagnetic indices with the
GPS positioning deviations. The storm period was
fragmented in sub-events, each characterised with
specific behavior of indices. Data interpretation and
subsequent analysis showed that positioning
deviations followed changes in geomagnetic field
intensity. Ionospheric TEC was introduced as an
entity which complemented the storm features and its
manifestation on positioning. Characteristic changes
in TEC behavior during times of geomagnetic
variations were observed. After the correlation
between all employed variables, regression tree
models were introduced to determine probable values
of positioning deviations as target variables. The
evaluation of models showed an acceptable level of
success, especially reflecting on height deviation
prediction as the most sensitive positioning
component. The proposed approach offers a simple
yet effective mean of assessment of GPS positioning
performance during space weather events directly or
indirectly affecting positioning deviations.
The storm was not classified as extreme; however,
it measurably influenced the positioning
performance. With ever growing GNSS utilisation,
such events could affect the simultaneous
performance of GNSS-based services that are
expected to be independent of each other. Further
research of similar events and employment of
additional indices is necessary to contribute to a more
comprehensive image of the solar-terrestrial
environment. Besides geomagnetic disturbances,
models should also employ solar, solar wind and
ionospheric indices, opening possibilities of predictive
357
analytics by monitoring upcoming disturbances from
their place of origin.
ACKNOWLEDGMENTS
This work has been fully supported by the University of
Rijeka under the project uniri-tehnic 18-66. The authors
would like to thank N. Ness, Principal Investigator of Bartol
Research Institute and to Dr. Howard Singer, NOAA Chief
Scientist. Authors appreciate and support the access to open
software tools: R, R Studio and Rattle, RTKLIB and Rinex-
GPS-TEC.
REFERENCES
Booker, H.G., Morphology of Ionospheric Storms,
Proceedings of the National Academy of Science of the United
States of America, PNAS, Washington DC, 1954, vol. 40,
no. 10, pp. 931–943.
Brčić, D., Ensuring sustainability through utilisation of
satellite navigation technology, Proceedings of the 2012
International Conference on Transport Sciences (ICTS);
Portoroz, Slovenia, Fakulteta za pomorstvo in promet,
2012, p 14.
Davies, K., Ionospheric Radio Propagation, National Bureau of
Standards, Washington DC, 1965.
Dyrud, L. et al., Ionospheric measurement with GPS:
Receiver techniques and methods. Radio Science, 2008,
vol. 43, no. 6, pp. 1-11.
Filić, M. and Filjar, R., Forecasting model of space weather-
driven GNSS positioning performance. Lambert Academic
Publishing, Saarbrucken, 2018.
Filjar, R., Brčić, D. and Kos, S., Single-frequency Horizontal
GPS Positioning Error response to a moderate
Ionospheric storm over Northern Adriatic, In: Weintrit,
A. (ed) Advances in Marine Navigation, Marine Navigation
and Safety of Sea Transportation, CRC Press, Boca Raton,
2013, pp. 49-56.
Goodman, J.M., Space Weather and Telecommunications,
Springer, New York, 2005.
Gurtner, W. and Estey, L., RINEX: The Receiver Independent
Exchange Format, V3.01, IGS Central Bureau, Pasadena,
2009, p. 44.
Hastie. T., Tibshirani, R. and Friedman, J., The Elements of
Statistical Learning, Springer, New York, 2009.
James, G., Witten, D., Hastie, T. and Tibshirani, R., An
Introduction to Statistical Learning, Springer, New York,
2009.
Kaplan, E.D. and Hegarty, C.J., Understanding GPS:
Principles and Application, 2nd ed., Artech House, Boston,
2006.
Kintner, P.M. and Ledvina, B., The ionosphere, radio
navigation and global navigation satellite systems,
Advances in Space Research, 2005, vol 35, no. 5, pp. 788-
811.
Klobuchar, J.A., Ionospheric corrections for timing
applications, Proceedings of the 20th Annual Precise Time
and Time Interval (PTTI) Application and Planning Meeting,
Naval Observatory, Washington DC, 1988, pp. 193-201.
Klobuchar, J.A., Ionospheric Effects on Earth-Space Propagation,
Environmental research paper No. 866, Air Force
Geophysics Laboratory, Hanscom, 1983, p 33.
Klobuchar, J.A., Ionospheric Time-Delay Algorithm for
Single-Frequency GPS Users, IEEE Transactions on
Aerospace and Electronic Systems, 1987, vol. 23, no.3, pp.
325–331.
Kouba, J.A., Guide to using International GNSS Service (IGS)
products, IGS Central Bureau, Pasadena, 2009, p. 34.
Lockwood, M., Wild, M.N., Stamper, R., Davis, C.J. and
Grande, M., Predicting Solar Disturbance Effects on
Navigation Systems, Journal of Navigation, 1999, vol. 52,
no. 2, pp. 203-216.
McMorrow, D., Impacts of Severe Space Weather on the Electric
Grid, Report JSR-11-320, The MITRE Corporation,
McLean, 2011, p. 107.
Mendillo, M., Storms in the ionosphere: Patterns and
processes for total electron content, Reviews of Geophysics,
2006, vol. 44, no. 4, p. 47.
Noll, C., The Crustal Dynamics Data Information System: A
resource to support scientific analysis using space
geodesy, Advances in Space Research, 2010, vol. 45, no. 12,
pp. 1421-1440.
Parkinson, B.W. and Spilker, Jr. J.J., Global Positioning
System: Theory and Applications, Vol. I, AIAA,
Washington DC, 1996.
Perrone, L. and De Franceschi, G., Solar, ionospheric and
geomagnetic indices, Annals of Geophysics, 1998, vol. 41,
no. 5-6, pp. 843-855.
Ross, W.J., The determination of ionospheric electron
content from satellite Doppler measurements: 2,
Experimental results, Journal of Geophysical Research,
1960, vol. 65, pp. 2607–2615.
Singer, H.J., Matheson, L., Grubb, R., Newman, A. and
Bouwer, S.D., Monitoring space weather with the GOES
magnetometers, Proc. SPIE 2812, GOES-8 and Beyond,
1996, p. 29.
Subirana, J. S., Zornoza, J. M. J. and Hernandez-Pajares, M.,
GNSS Data processing, Volume I: Fundamentals and
algorithms, ESA Communications, Noordwijk, 2013.
Thomas, M. et al., Global Navigation Space Systems: reliance
and vulnerabilities, RAENG, London, 2011.
Zolesi, B. and Cander, R. L. J, Ionospheric Prediction and
Forecasting, Springer, New York, 2014.