PreprintPDF Available

Machine learning-based multipath modelling in spatial-domain: a demonstration on GNSS short baseline processing

Authors:

Abstract and Figures

Multipath is the main unmodeled error source hindering high-precision Global Navigation Satellite System (GNSS) data processing. Conventional multipath mitigation methods, such as sidereal filtering (SF) and multipath hemispherical map (MHM), have certain disadvantages: they are either too complicated for implementation or not effective enough for multipath mitigation. In this study, we propose a machine learning (ML)-based multipath mitigation method. Multipath modelling was formulated as a regression task, and the multipath errors were fitted with respect to azimuth and elevation in the spatial-domain. We collected 30 days of 1 Hz GPS data to validate the proposed method. In total five short baselines were formed and multipath errors were extracted from the posfit residuals. ML-based multipath models, as well as observation-domain SF and MHM models, were constructed using 5 days of residuals before the target day and later applied for multipath correction. It was found that the XGBoost (XGB) method outperformed SF and MHM. It achieved the highest residual reduction rates, which were 24.9%, 36.2%, 25.5% and 20.4% for GPS P1, P2, L1 and L2 observations, respectively. After applying the XGB-based multipath corrections, kinematic positioning precisions of 1.6 mm, 1.9 mm and 4.5 mm could be achieved in east, north and up components, respectively, corresponding to 20.0%, 17.4% and 16.7% improvements compared to the original solutions. The effectiveness of the ML-based multipath model was further validated using 30 s sampling data. We conclude that the ML-based multipath mitigation method is effective, easy-to-use, and can be easily extended by adding auxiliary input features, such as signal-to-noise ratio (SNR), during model training.
Content may be subject to copyright.
Page 1/28
Machine learning-based multipath modelling in
spatial-domain: a demonstration on GNSS short
baseline processing
Yuanxin Pan ( yxpan@ethz.ch )
ETH Zurich
Gregor Möller
ETH Zurich
Benedikt Soja
ETH Zurich
Research Article
Keywords: GNSS, Multipath, Spatial-domain, Machine learning, XGBoost
Posted Date: February 9th, 2023
DOI: https://doi.org/10.21203/rs.3.rs-2555284/v1
License: This work is licensed under a Creative Commons Attribution 4.0 International License.
Read Full License
Page 2/28
Abstract
Multipath is the main unmodeled error source hindering high-precision Global Navigation Satellite System
(GNSS) data processing. Conventional multipath mitigation methods, such as sidereal ltering (SF) and
multipath hemispherical map (MHM), have certain disadvantages: they are either too complicated for
implementation or not effective enough for multipath mitigation. In this study, we propose a machine
learning (ML)-based multipath mitigation method. Multipath modelling was formulated as a regression
task, and the multipath errors were tted with respect to azimuth and elevation in the spatial-domain. We
collected 30 days of 1 Hz GPS data to validate the proposed method. In total ve short baselines were
formed and multipath errors were extracted from the post residuals. ML-based multipath models, as well
as observation-domain SF and MHM models, were constructed using 5 days of residuals before the
target day and later applied for multipath correction. It was found that the XGBoost (XGB) method
outperformed SF and MHM. It achieved the highest residual reduction rates, which were 24.9%, 36.2%,
25.5% and 20.4% for GPS P1, P2, L1 and L2 observations, respectively. After applying the XGB-based
multipath corrections, kinematic positioning precisions of 1.6 mm, 1.9 mm and 4.5 mm could be
achieved in east, north and up components, respectively, corresponding to 20.0%, 17.4% and 16.7%
improvements compared to the original solutions. The effectiveness of the ML-based multipath model
was further validated using 30 s sampling data. We conclude that the ML-based multipath mitigation
method is effective, easy-to-use, and can be easily extended by adding auxiliary input features, such as
signal-to-noise ratio (SNR), during model training.
1 Introduction
Global Navigation Satellite System (GNSS) has already become an essential part of our daily life and a
crucial part of the geodetic infrastructure (Rebischung et al., 2016). With the renement of error correction
models and the improvement of precise products provided by the International GNSS Service (IGS),
GNSS-based positioning precision can reach mm-level in static mode and cm-level in kinematic mode
(Bock et al., 2004; Kouba, 2015; Choy et al., 2016). However, multipath still remains the main unmodelled
error source due to its nonlinear nature. It degrades the contribution of GNSS to applications demanding
high precision, such as earthquake early warning (Larson, 2009).
Multipath is the effect of simultaneous reception of direct and reected GNSS signals. It is almost
inevitable due to the nondirectional nature of GNSS antennas. Apart from choosing a less reective
environment, hardware- and software-based measures are usually adopted to reduce multipath. The
hardware-based methods can be divided into antenna design and receiver improvement, such as choke
ring and narrow-band correlation (Van Dierendonck et al., 1992; McGraw et al., 2004). However, they can
only reduce part of the multipath error (Park et al., 2004). The software-based approaches include various
ltering methods utilizing the frequency signature of multipath (Satirapod and Rizos, 2005). However, it is
hard to apply such ltering when the multipath frequency range overlaps with that of signals of interest.
Signal-to-noise ratio (SNR) measured by GNSS receivers can be used for multipath characterization or
observation weighting (Bilich et al., 2008; Su et al., 2021). However, its eciency is dependent on the SNR
Page 3/28
data quality and antenna gain pattern. Sidereal ltering (SF) is a widely used method to mitigate
multipath for high-precision GNSS data processing (Genrich and Bock, 1992; Bock et al., 2004). The idea
is that the geometric relation between the Global Positioning System (GPS) constellation and a static
station will repeat every sidereal day, and the positioning error induced by multipath will also repeat after
the same period. Hence, the coordinate time series of previous days, with proper time shifts, can be used
to correct the multipath for the target day. The key to implement SF is to calculate the correct orbit repeat
period for each satellite, since the actual orbit repeat period of GPS is not exactly one sidereal day and
even varies with different satellites (Choi et al., 2004; Agnew and Larson, 2006). When it comes to multi-
GNSS, the case is more complicated and SF can no longer be applied in the coordinate-domain. In order
to solve this problem and use the individual orbit repeat period for each satellite, observation-domain SF
was rst proposed by Zhong et al. (2010) for baseline processing and was also successfully applied to
precise point positioning (PPP) and multi-GNSS processing (Atkins and Ziebart, 2015; Ye et al., 2014;
Geng et al., 2018). It extracts multipath corrections from postt residuals of previous days and applies
them to the observations of each satellite on the target day after shifting the corrections by individual
orbit repeat periods. Although SF can effectively mitigate multipath errors, it is cumbersome to implement
due to the different orbit repeat periods of GNSS satellites, and it is less effective for observations of low
sampling rate.
The spatiotemporal repeatability of multipath can also be modelled in the spatial domain. It is based on
the fact that multipath errors mainly depend on satellite positions in a skyplot, and thus a multipath
correction model can be established with respect to azimuth and elevation angles in a topocentric
coordinate system. Cohen and Parkinson (1991) proposed a multipath lookup table to model the
reective environment around the station. Fuhrmann et al. (2014) used congruent grids with similar
shapes and sizes to generate multipath maps. Dong et al. (2015) named this kind of spatial domain-
based multipath model as multipath hemispherical map (MHM) and compared its performance with SF
using 1 Hz GPS data from a dual-antenna receiver. It was concluded that similar multipath mitigation
performance could be achieved with both methods but MHM was less effective for high-frequency
multipath. However, MHM is satellite independent and is easy to implement and use (Zheng et al., 2019;
Lu et al., 2021). Wang et al. (2019) modied the MHM method by introducing a set of trend surface
coecients for each grid to capture the multipath variation within a grid. It was found that the modied
MHM method achieved about 5% more residual reduction rate than MHM, but it complicated the
application of the original MHM method.
Over the last decade, articial intelligence, especially machine learning (ML), has become more and more
prominent in geosciences (Li et al., 2011; Beroza et al., 2021; Crocetti et al., 2021; Aichinger et al., 2022).
Such data-driven algorithms are suitable for solving nonlinear problems, including classication and
regression tasks. ML algorithms have already been applied to GNSS multipath and non-line-of-sight
(NLOS) signal classication. It was shown that 75%-90% classication accuracy could be achieved with a
support vector machine (SVM) when appropriate input features were used (Hsu, 2017; Xu et al., 2020).
Suzuki et al. (2020) trained a convolutional neural network (CNN) to detect NLOS signals based on the
output of multiple GNSS signal correlators of a software-dened receiver and reported a 98%
Page 4/28
classication accuracy. Li et al. (2022) demonstrated the advantage of deep neural network (DNN)-based
signal correlation schemes in a receiver tracking loop over standard correlation schemes regarding
multipath mitigation. Tao et al. (2020, 2021) used neural networks to mine the multipath features in
coordinate and frequency domains, respectively, and reported better multipath mitigation performance
than conventional methods. However, currently there is no research that studies the possibility of
multipath modelling in the spatial-domain with ML.
The focus of this paper is to investigate the potential of ML algorithms on multipath modelling in the
spatial-domain. We formulate multipath modelling as a regression task for ML algorithms. The multipath
errors are tted with respect to azimuth and elevation angles in the skyplot. The benet of SNR
measurements for multipath modelling is also examined. Three widely used ML methods, i.e., random
forest (RF), extreme gradient boosting (XGB) and multilayer perceptron (MLP), are tested regarding
multipath mitigation for short baselines. The remainder of this paper is organized as follows: principles
of multipath modelling are introduced in Section 2. The data used in this study is described in Section 3.
The ML-based multipath mitigation results are displayed and discussed in Section 4. Conclusions and
outlooks are given in Section 5.
2 Multipath Modelling
Multipath is the composite of direct and reected GNSS signals, and it cannot be modelled thoroughly
due to its nonlinearity. But under the assumption of specular reection, the multipath errors of
pseudorange and carrier phase can be modelled as (Bilich et al., 2007):
1
where and are multipath errors of pseudorange and carrier phase, respectively. We denote the
reection coecient as , which is the amplitude ratio between the reected signal and the direct signal.
The geometric path delay is denoted as , and is the phase offset of the reected signal, which is
caused by the extra path delay and phase shift due to the reection.
The SNR measured by a receiver is a useful indicator of multipath errors. It contains the reection
information of the environment:
2
PMP
=
Φ
MP
= tan1
α
δ
cos
φ
1 +
α
cos
φ
α
sin
φ
1 +
α
cos
φ
PMP
Φ
MP αδ φ
SNR
2=
A
2
d
+
A
2
m
+ 2
AdAm
cos
φ
Page 5/28
where and are the amplitudes of direct and reected GNSS signals. The symbol has the same
meaning as in Eq.(1). It can be noted that is the common underlying parameter for multipath and SNR,
which means that SNR measurements could be benecial for multipath modelling.
Multipath extraction
GPS data is processed in baseline mode to extract multipath errors. The mathematical models for short
baseline processing can be formulated as follows:
5
where is the double-differencing operator, and are pseudorange and carrier phase
measurements on frequency , respectively, is the geometric distance between the receiver and the
satellite, is the signal wavelength, and is the integer ambiguity. Pseudorange and carrier phase
multipath errors are denoted as and , respectively. The observation noises of pseudorange and
carrier phase are denoted as and , respectively. Receiver and satellite clocks are eliminated by
double-differencing. Troposphere and ionosphere delays can be neglected in short baseline processing.
Earth tides, phase windup and satellite antenna PCO/PCV (Phase Center Offset/Variation) are not
necessary to be corrected for short baselines. Receiver antenna PCO/PCV can also be neglected if the
same type of antenna is used.
The baseline data is usually processed in static mode to squeeze all the multipath errors into posterior
residuals. Then these double-differenced (DD) residuals are converted to single-differenced (SD) ones by
adding a zero-mean hypothesis for each epoch (Alber et al., 2000). Otherwise, each DD residual is related
to two satellites with different azimuth and elevation angles, which is not suitable for multipath model
construction. The converted SD residuals are further low-pass ltered with an empirical corner frequency
of 0.1 Hz to remove the observation noise (Choi et al., 2004; Geng et al., 2018). These ltered residuals
can be predominantly treated as multipath errors and will be used for multipath modelling.
ML-based multipath modelling
ML algorithms have been proven to be powerful tools for regression tasks. The main advantage of ML
over classic spatial interpolation algorithms is that not only azimuth and elevation angles but also
auxiliary information, such as SNR measurements, can be utilized for interpolation. There are a lot of ML
algorithms suitable for regression tasks. Among them, ensemble learning algorithms, including bagging
and boosting, are usually ranked among the best-performing methods. Besides, articial neural networks
(ANN) are also commonly used ML algorithms. Hence, three representative ML algorithms, including
random forest (RF), extreme gradient boosting (XGB) and multilayer perceptron (MLP) are selected as the
candidate methods for multipath modelling in this study. RF is an ensemble learning method that outputs
the average results of a set of randomized decision trees (Breiman, 2001). It can overcome the overtting
AdAmφ
φ
{
Δ
Pi
= Δ
ρ
+ Δ
Mi
+ Δ
εi
Δ
Li
= Δ
ρ
+
λi
Δ
Ni
+ Δ
mi
+ Δ
σi
Δ
PiLi
fiρ
λiNi
Mimi
εiσi
Page 6/28
issue of a single decision tree and usually can achieve high accuracy without complex conguration.
XGB is an open-sourced gradient boosting framework (Friedman, 2001; Chen and Guestrin, 2016). The
basic idea is that a set of decision trees are trained sequentially to better t the samples with larger
residuals, and it is widely used due to its high performance. MLP is a type of ANN with fully connected
nodes. It consists of three parts, i.e., input layer, hidden layer(s) and output layer. Nonlinear activation
functions are used at each node, and thus can simulate the nonlinear relation between input and output.
Data preparation is the key to ML model training. Two sets of data, i.e., input features and a target vector
need to be provided. Apart from azimuth and elevation angles, SNR is also tested as an additional input
feature for model training in this study. The SD residuals extracted from baseline processing form the
target vector. ML models are trained to best t the relation between input features and the target vector.
Note that all the training data should be cleaned for outliers and normalized to improve training stability
and model performance. The basic procedures of data processing and ML-based multipath mitigation
are illustrated in Fig.1. More details about optimal ML model construction can be found in Section 4.2.
The observation-domain SF and MHM multipath models are built on the same data set for performance
comparison. Note that observation-domain SF is simply denoted as SF in the remaining text for clarity.
Specically, the actual orbit repeat period required by SF is calculated using broadcast ephemerides for
each satellite (Choi et al., 2004), and the MHM model is constructed with 1° by 1° grids for improved
model stability and effectiveness (Dong et al., 2015). Multipath models for GPS P1, P2, L1 and L2
observations are constructed individually. Although pseudorange multipath can be quantied with the so-
called multipath combination if dual-frequency measurements are available, it is still meaningful to
investigate pseudorange multipath modelling since the multipath model can be applied for single-
frequency data.
3 Data
We obtained 30 days (DOY 244–273 of 2021) of 1 Hz high-rate GPS data from Curtin University to test
the ML-based multipath mitigation method. There were four GPS antennas on the rooftop of Curtin
University (Fig.2). The antenna connected to a Trimble NetR9 receiver (CUT0) was used as the reference
station. The other three antennas were all connected to two different receivers, forming six rover stations,
i.e., CUAA/CUTA, CUBB/CUTB and CUCC/CUTC (see details in Table1). Here, a station denotes the
combination of an antenna and a receiver. Station CUBB was excluded in the following studies since
there were many gaps in its data. Hence, ve short baselines were formed between CUT0 and rover
stations for multipath mitigation experiments. The baselines were processed with a modied version of
the RTKLIB software (Takasu, 2009). The coordinates of the reference station were xed to a mean SPP
solution (~ 5 m precision). Uncombined observations of pseudorange and carrier phase on dual
frequencies were utilized for parameter estimation. Since the distance between the reference station and
any rover station was less than 10 m, tropospheric and ionosphere delays were eliminated by differencing
between stations. Only rover positions and DD ambiguities were estimated in a Kalman lter. It was worth
noting that all the stations had the same type of antenna (Table1), which meant receiver PCV errors
Page 7/28
could be eliminated through differencing and would not affect multipath modelling. Finally, multipath
errors were extracted from the postt residuals according to the method described in Section 2 and were
used for multipath modelling in the following experiments.
Table 1
Detailed congurations of GPS stations used in this study
Station Receiver type Firmware version Antenna type
CUT0 Trimble NetR9 5.45 TRM59800.00 SCIS
(Choke ring antenna)
CUTA 5.22
CUTB 5.22
CUTC 5.45
CUAA Javad TRE_G3T DELTA 3.7.9
CUBB
CUCC
4 Results
We rst analyze the characteristics of the extracted multipath errors. Then, different input features and
ML algorithms are explored to establish the multipath models. Finally, the ML-based multipath mitigation
method is compared to the conventional observation-domain SF and MHM methods regarding residual
reduction and positioning improvement. Considering the similar data quality and multipath environment,
we take the station CUCC as the example for specic analysis, and present the statistical results for all
the stations.
4.1 Multipath characteristics analysis
The key to set up a reliable multipath model is the spatiotemporal repeatability of multipath. Figure3
shows the low-pass ltered residuals of satellite G04 at CUCC station on DOY 244, 245 and 249. Note
that the residual time series of DOY 245 and 249 are shifted by corresponding orbit repeat times to better
show the correlation with the residuals of DOY 244. The Pearson correlation coecients between DOY
244 and 249 can reach 0.68, 0.70, 0.82 and 0.86 for P1, P2, L1 and L2, respectively. It indicates that the
environment around the stations is quite stable during the experiment periods, and it also conrms that
the zero-mean constraint for residual conversion is effective according to the repeatability characteristics
of multipath. Considering the good temporal correlation, we stacked 5 days of residuals before the target
day to enhance the multipath signals during modelling (Dong et al., 2016; Wang et al., 2019). We also
checked the multipath correlation between both frequencies for pseudorange and carrier phase,
respectively. The correlation coecients between P1 and P2 residuals are below 0.2 and most of the time
close to 0. This indicates the multipath errors of different pseudorange measurements are not correlated.
Page 8/28
For carrier phase residuals on dual-frequency, the correlation coecients vary between 0 and 0.5. Hence,
there is no denite relation between L1 and L2 multipath effects. Considering the same path delay but
different wavelengths of L1 and L2 carriers, the phase of the two reected carriers is usually
unsynchronized, and thus the disturbance on the direct carrier signals will be different and uncorrelated
(Bilich et al., 2007).
4.2 Optimal ML-based multipath model setup
To select the best ML algorithms and input feature combinations, we used 30 days of data from the
CUCC station to evaluate the performance of each combination regarding the residual reduction rate. The
three candidate ML algorithms included RF, XGB and MLP, and the potential input features included
azimuth, elevation and SNR. Since azimuth and elevation were necessary for spatial interpolation, there
were only two choices for input features, i.e., with or without SNR. ML-based multipath models with the
six possible combinations of algorithms and input features were trained on the residuals of DOY 244–
248 and validated using the residuals of DOY 249, respectively. Grid search was adopted for
hyperparameter tuning. The optimal hyperparameters (Table2) were determined based on the residual
reduction rates for DOY 249. Note that adding SNR as an additional input feature had little impact on the
optimal hyperparameters for each ML algorithm (i.e., RF, XGB or MLP) according to our experiments. After
determining the best hyperparameters, model training and testing were repeated for the remaining 24
days (see Fig.1) and mean multipath reduction rates were calculated for each combination. This meant 5
days of residuals before the target day were used to train the model and later it was applied for multipath
mitigation for the target day.
Figure4 shows the average residual reduction rates for all six combinations. It indicates that including
SNR as an additional feature does not help to improve the model performance, especially for RF and
XGB. The reason might be that the numerical precision of SNR values in the RINEX (Receiver Independent
Exchange Format) les is not high enough. The Trimble and Javad receivers involved in this study only
record SNR measurements with a precision of 0.2 dB-Hz and 0.25 dB-Hz, respectively. Such coarse SNR
increments are not precise enough to improve multipath modelling. Bilich et al. (2007) also found this
issue and reported that it was receiver model dependent. Still, adding SNR as an additional feature for
MLP can improve its performance for pseudorange multipath mitigation. Using only azimuth and
elevation angles as input features is sucient for RF and XGB models. Residual reduction rates of 25%,
36%, 30% and 25% can be achieved for P1, P2, L1 and L2, respectively. It is worth pointing out that
although RF can conduct multivariate regression (i.e., one multipath model for four observables), no
obvious improvement can be observed compared to the results of building multipath models individually.
Since XGB with azimuth and elevation as input features can achieve highest residual reduction rates, we
only present the results of this combination for ML-based methods in the following experiments.
Page 9/28
Table 2
Optimal set of hyperparameters derived from grid search
Algorithm Hypeparameter Value
RF n_estimators 10
max_depth 30
criterion squared_error
XGB n_estimators 40
max_depth 20
criterion squared_error
MLP hidden_layer_sizes (128,128,128,128,128,128,128)
activation relu
solver adam
4.3 Multipath mitigation test
After picking the optimal ML algorithm and input features, we tested and compared the multipath
mitigation performance for three different methods: SF, MHM and XGB. The improvements in kinematic
relative positioning were also evaluated.
4.3.1 Multipath model
The multipath models based on the XGB method are visualized in Fig.5 for station CUCC on DOY 251. It
can be seen that most severe multipath errors concentrate in the low elevation areas, and there is no
obvious pattern difference with respect to azimuth. This is because there is no strong reection source
around the station, and most reected signals come from the surrounding grounds (Fig.2). Such an
observation environment is similar to most IGS stations and can make the conclusions of this study
generally applicable. We further plotted the multipath correction time series of SF, MHM and XGB in Fig.6
(a) to directly compare their capability of modelling multipath. The corresponding low-pass ltered
residuals of satellite G10 are also included as the reference. It can be found that the multipath models of
SF and XGB are in good agreement with the low-pass ltered residuals for both pseudorange and carrier
phase on dual frequencies. They successfully replicate both the long- and short-term variations induced
by multipath. However, the multipath model from MHM can only capture the long-term tendency but not
the short-term changes, i.e., high-frequency components. This drawback is most obvious during the
period from 14 h to 16 h when the satellite is at low elevation and multipath changes fast. The MHM
multipath model in this period resembles a low-resolution version of SF and XGB models.
Figure6 (b) exhibits the power spectral density (PSD) for L1 multipath models generated by three
methods as well as the low-pass ltered residuals. The MHM model has a lower power density between
Page 10/28
the frequency range from 17s to 60 s. Compared to XGB, MHM is about 7 dB lower for the high-frequency
multipath components, which accounts for around 80.5% lower signal power. This explains the low
resolution of the MHM model in Fig.6 (a). For lower frequencies, MHM agrees well with the other two
methods. In contrast, the PSDs of SF and XGB are in good agreement with that of low-pass ltered
residuals from 17s to the lowest frequency. The PSD drop before 0.1 Hz of the SF model and the low-
pass ltered residuals is caused by the low-pass lter used to remove the white noise in raw residuals.
The higher noise level in the XGB model between 2 s and 17 s is most probably caused by spatial
interpolation errors, but it does not affect the multipath mitigation effect since the noise magnitude is
very small compared to multipath errors. A similar phenomenon is also observed for MHM, although it is
not expected since the MHM model averages the residuals within each grid and the noise level should be
lower. Hence, the higher noise level is possibly an artifact caused by the step signals generated by the
low-resolution MHM model as shown in Fig.6 (a). Overall, the PSD analysis conrms that the ML-based
multipath model can achieve similar performance as SF and outperforms the MHM model due to the
advantage of spatial interpolation.
4.3.2 Residual reduction
Figure7 shows the posterior residuals of G04 at station CUCC on DOY 251 and those corrected using SF,
MHM and XGB methods. The multipath errors are effectively reduced by all three methods for
pseudorange and carrier phase on dual frequencies, especially for periods when the satellite is at low
elevations. The RMS of the residuals corrected with XGB is the smallest among the three methods,
reaching 0.55 m, 0.27 m, 2.25 mm and 2.52 mm for P1, P2, L1 and L2, respectively. Compared to the raw
residuals, the improvements are 26.7%, 41.3%, 36.6% and 30.6%, respectively. Here, it can be found that
the reduction rate for P1 is much smaller than for P2. This can be explained by the higher noise level of
P1 residuals. The multipath mitigation performance of SF is similar to that of XGB, and the RMS
differences between them are only 0.01 m, 0.01 m, 0.00 mm and 0.02 mm for P1, P2, L1 and L2,
respectively. In contrast, MHM can only achieve improvements of 10.6%, 26.1%, 25.1% and 25.3% for the
four observables. That is because the MHM method models multipath using 1° by 1° grids, and it cannot
capture the high-frequency multipath components. This can be seen in the L1 residual time series
between 14 h and 16 h in Fig.7. There are still obvious uctuations in this period after being corrected
with MHM, especially the variations near 15 h. Such uctuations nearly disappear when XGB or SF
models are applied.
The mean residual reduction rates using SF, MHM and XGB over all ve stations and 25 days are
displayed in Fig.8. Overall, the results are consistent with those shown in Fig.7, i.e., XGB performs
similarly to SF and better than MHM regarding residual reduction rates. After multipath mitigated with
XGB, RMS improvements of 24.9%, 36.2%, 25.5% and 20.4% can be achieved for P1, P2, L1 and L2
residuals, respectively. The reduction is 0.1–0.7% and 2.0–2.8% larger than for SF for carrier phase and
pseudorange residuals, respectively. In contrast, the reduction rates achieved with MHM are 13.7%, 14.3%,
8.0% and 3.5% less than for XGB for P1, P2, L1 and L2, respectively, due to its deciency of modelling
high-frequency multipath signals.
Page 11/28
Up to now, we have always used the residuals of the latest 5 days before the target day to set up the
multipath model. Although the best correction effect can be obtained in this way, updating the model in a
daily manner is cumbersome. It would be benecial, especially for real-time applications, if the multipath
model can be applied to subsequent days without too much precision loss. Hence, we tested the model
validity period for all ve stations and compare the performance among SF, MHM and XGB. Residuals
from DOY 244–248 are used to set up the multipath model, and later it is applied for multipath correction
on days from DOY 249 to 273. The mean residual reduction rates over ve stations on each day are
plotted in Fig.9. It can be found that the multipath correction effect of the three different methods
gradually degrades over the whole test period. XGB achieves the highest residual reduction rates for all
four observables on the 25 test days. The reduction rates using XGB drop from 25.2%, 35.6%, 25.1% and
19.4–9.0%, 17.7%, 16.1% and 14.4% for P1, P2, L1 and L2 residuals, respectively. It means the multipath
correction effect on DOY 273 is only half of that on DOY 249. A 5–7 days update rate seems to be a good
trade-off between model validity and the workload of data processing. In this circumstance, it can still
achieve 90% multipath correction effect of the daily updated model. SF performs similarly to XGB on the
rst day, but its performance rapidly drops on the subsequent three days and then decreases at a linear
pace. That is mainly because the effectiveness of the SF model heavily depends on the accurate orbit
repeat time for each satellite on each day and any deviations between the computed and real orbit repeat
time will impact the model performance. The MHM model performance is more stable, with the reduction
rates dropping by 5.9%, 8.7%, 6.3% and 5.6% over the test periods for P1, P2, L1 and L2, respectively. It
conrms that MHM model mainly captures the lower frequency multipath signals as they are more stable
over time.
4.3.3 Positioning improvement
We further applied the three different multipath models for kinematic relative positioning. The model
performance was evaluated regarding the positioning precision improvement compared to the solutions
without correction. The daily static coordinates were used as the benchmark for calculation of
positioning RMS. Note that multipath correction was not applied for static solutions as the impact of
multipath on daily static positioning could be neglected. We nally obtained 125 time series (at ve
stations over 25 days) for each type of solution, i.e., raw (without correction), corrected with SF, MHM and
XGB.
Figure10 depicts the displacements for four different solutions at station CUCC on DOY 273. The raw
solution contains many variations induced by multipath spanning from tens of seconds to half an hour,
which are evident in all three coordinate components. These variations are effectively mitigated by
applying the multipath models of SF, MHM and XGB. The RMS of the solution corrected by XGB are 1.4
mm, 1.9 mm and 4.4 mm for east, north and up components, respectively, which are equal to those of SF.
It is interesting that the MHM model can reach comparable positioning precisions, especially considering
its disadvantage to capture high-frequency multipath signals. The RMS values are only 0.1 mm, 0.1 mm
and 0.2 mm larger than the other two models in east, north and up components, respectively. Usually, the
high-frequency multipath occurs when a satellite is at low elevations. Such low-elevation observations
Page 12/28
are down-weighted during data processing. This can explain the reasonable positioning precision of
MHM although it is decient in high-frequency multipath modelling. The mean positioning precisions
over all ve stations and 25 days are listed in Table3. Again, XGB and SF can achieve the highest
precisions, which are 1.6 mm, 1.9 mm and 4.5 mm for east, north and up components, respectively.
Compared to the raw solutions, the improvements are about 20.0%, 17.4% and 16.7% for the three
components. The performance of MHM is a bit worse compared to XGB and SF, but it can still reach
15.0%, 13.0% and 13.0% precision improvements for east, north and up components, respectively.
Table 3
Mean RMS of 1 Hz displacements in east, north and
up components for 4 types of solutions over all 5
stations and 25 days
Method Kinematic positioning precision (mm)
East North Up
Raw 2.0 2.3 5.4
Sidereal 1.6 1.9 4.5
MHM 1.7 2.0 4.7
XGB 1.6 1.9 4.5
4.4 Multipath mitigation for 30 s sampling data
In the last section, we have demonstrated the multipath mitigation performance of the XGB model using
1 s GPS data. However, 30 s is the more common sampling rate for most geodetic stations, such as the
IGS network. Hence, we further validate the XGB model using GPS data of 30 s interval.
We reprocessed the data at 30 s sampling rate for all ve stations and utilized the 30 s sampling
residuals from 5 days before each target day to set up the multipath models for SF, MHM and XGB,
respectively. Then the residuals of P1, P2, L1 and L2 are corrected with the corresponding models. The
daily residual reduction rates are plotted in Fig.11 and the mean values over 25 days are given in Table4.
We nd that the reduction rates of XGB are the highest among the three different methods. The
performance of XGB is almost constantly 1.9% and 2.6% higher than for SF and MHM for P1 residuals
and 2.7% and 4.5% for P2. For carrier phase residuals, the reduction rates of XGB are about 1.0% and
2.4% higher than for SF and MHM for L1, and 0.4% and 1.3% for L2. This demonstrates the superiority of
ML methods for multipath mitigation. In the circumstance of 30 s sampling rate, the error of the orbit
repeat time calculation might be up to 15 s, which will degrade the SF model performance. For 30 s data,
the frequency of multipath signals will not be higher than 60 s. Hence, the performance of the MHM
model is closer to SF and XGB. But there will also be fewer data points within each grid cell, which might
degrade the stability of the MHM model. The corresponding kinematic positioning results are listed in
Table5. It is found that XGB can achieve 0.1 mm higher precision in both north and up components than
Page 13/28
those of SF and MHM. Compared to the raw solutions, precision improvements of 15.0%, 13.0% and
11.3% can be achieved with the XGB model in east, north and up components, respectively.
Table 4
Mean residual reduction rates for 30 s
data over all ve stations and 25 days
Model Residual reduction rate (%)
P1 P2 L1 L2
SF 10.2 21.8 16.3 15.6
MHM 9.5 20.0 14.9 14.7
XGB 12.1 24.5 17.3 16.0
Table 5
Mean RMS of 30 s displacements in east, north and
up components for four types of solutions over all
ve stations and 25 days
Method Kinematic positioning precision (mm)
East North Up
Raw 2.0 2.3 5.3
Sidereal 1.7 2.1 4.8
MHM 1.7 2.1 4.8
XGB 1.7 2.0 4.7
5 Conclusion
Multipath is the main unmodelled error in high-precision GNSS data processing. It hinders the
achievement of mm-level kinematic positioning precision, which especially impacts the application of
GNSS for seismology studies and structure health monitoring. In this study, we proposed an ML-based
multipath mitigation method. It takes azimuth and elevation as input features and outputs multipath
corrections for pseudorange and carrier phase on both frequencies. Owing to its ability of spatial
interpolation, it can overcome the shortcoming of the conventional MHM method, i.e., deciency in
capturing high-frequency multipath signals. With 30 days of 1 Hz GPS data from ve baselines on the
rooftop of Curtin University, we validated the superiority of the ML-based multipath mitigation method
over conventional methods. The best ML algorithm (XGB) and optimal input features (azimuth and
elevation angles) are selected based on the residual variance reduction rate. However, SNR
measurements cannot improve the model performance and it might be attributed to the insucient
numeric precision. The conventional SF and MHM models were used for comparison. We demonstrate
that the XGB model can achieve 24.9%, 36.2%, 25.5% and 20.4% reduction rates for P1, P2, L1 and L2
Page 14/28
residuals, respectively. Such performance is similar to SF but without the inconvenience of computing the
orbit repeat period for each satellite. XGB can reach 14.0% and 5.8% more reduction rates than MHM for
pseudorange and carrier phase residuals, respectively. After applying the XGB model, kinematic
positioning precisions of 1.6 mm, 1.9 mm and 4.5 mm can be achieved in east, north and up
components, respectively, which are 20.0%, 17.4% and 16.7% improvements compared to the raw
solutions. The effectiveness of the XGB model for 30 s sampling data was also evaluated and compared
to that of SF and MHM. It conrms that the advantage of spatial interpolation still holds for low-sampling
data. Residual reduction rates of 12.1%, 24.5%, 17.3% and 16.0% can be reached for P1, P2, L1 and L2,
respectively, which are better than for SF and MHM.
Although we only demonstrated the ML-based multipath mitigation using baseline data in this study, it is
also valid for PPP multipath modelling and mitigation according to our preliminary internal tests. Since
the ML-based model has the merit of ease of use, it is also promising for real-time applications, such as
structural health monitoring. Besides, the ML-based model can be extended to include additional input
features, such as environmental information and SNR measurements with higher numerical precision.
This might be helpful to further improve the model, especially in long-term performance. Finally, more
sophisticated ML algorithms are worth further investigation. As demonstrated in this study, basic tree-
based ML algorithms can perform well for multipath modelling. More powerful algorithms, such as deep
learning and reinforcement learning, should be further examined for multipath mitigation in future
studies.
Declarations
Ethics approval and consent to participate: Not applicable.
Consent for publication: All authors approved the manuscript for publication.
Availability of data and materials: The high-rate GNSS data used in this study is available at
http://saegnss2.curtin.edu.au/ldc/.
Competing interests: The authors declare no conict of interest.
Funding: This research received no external funding.
Authors contributions: YP designed the study, analyzed the data and wrote the paper. GM and BS
supervised the study and revised the manuscript. All authors reviewed the manuscript and approved it for
publication.
Acknowledgements: The authors would like to thank Curtin GNSS-SPAN Group for the access to the high-
rate GNSS data and Amir Allahvirdi-Zadeh for providing the station photos.
References
Page 15/28
1. Agnew, D. C., & Larson, K. M. (2006). Finding the repeat times of the GPS constellation. GPS
Solutions, 11(1), 71-76. https://doi.org/10.1007/s10291-006-0038-4
2. Aichinger-Rosenberger, M., Brockmann, E., Crocetti, L., Soja, B., & Moeller, G. (2022). Machine learning-
based prediction of Alpine foehn events using GNSS troposphere products: rst results for Altdorf,
Switzerland. Atmospheric Measurement Techniques, 15(19), 5821-5839.
3. Alber, C., Ware, R., Rocken, C., & Braun, J. (2000). Obtaining single path phase delays from GPS
double differences. Geophysical Research Letters, 27(17), 2661-2664.
https://doi.org/10.1029/2000GL011525
4. Atkins, C., & Ziebart, M. (2015). Effectiveness of observation-domain sidereal ltering for GPS precise
point positioning. GPS Solutions, 20(1), 111-122. https://doi.org/10.1007/s10291-015-0473-1
5. Beroza, G. C., Segou, M., & Mostafa Mousavi, S. (2021). Machine learning and earthquake
forecasting—next steps. Nature communications, 12(1), 1-3. https://doi.org/10.1038/s41467-021-
24952-6
. Bilich, A., Axelrad, P., & Larson, K. M. (2007). Scientic Utility of the Signal-to-Noise Ratio (SNR)
Reported by Geodetic GPS Receivers. Proceedings of the 20th International Technical Meeting of the
Satellite Division of The Institute of Navigation (ION GNSS 2007), Fort Worth, TX.
7. Bilich, A., Larson, K. M., & Axelrad, P. (2008). Modeling GPS phase multipath with SNR: Case study
from the Salar de Uyuni, Boliva. Journal of Geophysical Research, 113(B4).
https://doi.org/10.1029/2007jb005194
. Bock, Y., Prawirodirdjo, L., and Melbourne, T. I. (2004). Detection of arbitrarily large dynamic ground
motions with a dense high-rate gps network. Geophysical Research Letters, 31(6).
https://doi.org/10.1029/2003GL019150
9. Breiman, L. (2001). Random forests. Machine learning, 45(1), 5-32.
https://doi.org/10.1023/A:1010933404324
10. Chen, T., & Guestrin, C. (2016). Xgboost: A scalable tree boosting system. In Proceedings of the 22nd
acm sigkdd international conference on knowledge discovery and data mining, pp. 785-794.
https://doi.org/10.1145/2939672.2939785
11. Choi, K., Bilich, A., Larson, K. M., & Axelrad, P. (2004). Modied sidereal ltering: Implications for high-
rate GPS positioning. Geophysical Research Letters, 31(22). https://doi.org/10.1029/2004gl021621
12. Choy, S., Bisnath, S., & Rizos, C. (2016). Uncovering common misconceptions in GNSS Precise Point
Positioning and its future prospect. GPS Solutions, 21(1), 13-22. https://doi.org/10.1007/s10291-
016-0545-x
13. Cohen C, Parkinson B (1991) Mitigating multipath error in GPS-based attitude determination.
Advances in the astronautical sciences, AAS guidance and control conference, Keystone. Univelt, San
Diego, pp 74–78
14. Crocetti L, Schartner M, Soja B. (2021). Discontinuity Detection in GNSS Station Coordinate Time
Series Using Machine Learning. Remote Sensing, 13(19):3906. https://doi.org/10.3390/rs13193906
Page 16/28
15. Dong, D., Wang, M., Chen, W., Zeng, Z., Song, L., Zhang, Q., Cai, M., Cheng, Y., & Lv, J. (2015).
Mitigation of multipath effect in GNSS short baseline positioning by the multipath hemispherical
map. Journal of Geodesy, 90(3), 255-262. https://doi.org/10.1007/s00190-015-0870-9
1. Friedman, J. H. (2001). Greedy function approximation: a gradient boosting machine. Annals of
statistics, 1189-1232. https://doi.org/10.1214/aos/1013203451
17. Fuhrmann, T., Luo, X., Knöper, A., & Mayer, M. (2014). Generating statistically robust multipath
stacking maps using congruent cells. GPS Solutions, 19(1), 83-92. https://doi.org/10.1007/s10291-
014-0367-7
1. Geng, J., Pan, Y., Li, X., Guo, J., Liu, J., Chen, X., & Zhang, Y. (2018). Noise Characteristics of High-Rate
Multi-GNSS for Subdaily Crustal Deformation Monitoring. Journal of Geophysical Research: Solid
Earth, 123(2), 1987-2002. https://doi.org/10.1002/2018jb015527
19. Genrich, J. F. and Bock, Y. (1992). Rapid resolution of crustal motion at short ranges with the global
positioning system. Journal of Geophysical Research: Solid Earth, 97(B3):3261–3269.
https://doi.org/10.1029/91JB02997
20. Hsu, L. T. (2017). GNSS multipath detection using a machine learning approach. In 2017 IEEE 20th
International Conference on Intelligent Transportation Systems (ITSC) (pp. 1-6).
21. Kouba, J., (2015). A guide to using International GNSS Service (IGS) products.
http://kb.igs.org/hc/en-us/article_attachments/203088448/UsingIGSProductsVer21_cor.pdf
22. Larson, K. M. (2009). GPS seismology. Journal of Geodesy, 83(3-4), 227-233.
https://doi.org/10.1007/s00190-008-0233-x
23. Li, H., Borhani-Darian, P., Wu, P., & Closas, P. (2022). Deep Neural Network Correlators for GNSS
Multipath Mitigation. IEEE Transactions on Aerospace and Electronic Systems, 1-23.
https://doi.org/10.1109/taes.2022.3197098
24. Li, J., Heap, A. D., Potter, A., & Daniell, J. J. (2011). Application of machine learning methods to
spatial interpolation of environmental variables. Environmental Modelling & Software, 26(12), 1647-
1659. https://doi.org/10.1016/j.envsoft.2011.07.004
25. Lu, R., Chen, W., Dong, D., Wang, Z., Zhang, C., Peng, Y., & Yu, C. (2021). Multipath mitigation in GNSS
precise point positioning based on trend-surface analysis and multipath hemispherical map. GPS
Solutions, 25(3). https://doi.org/10.1007/s10291-021-01156-5
2. McGraw, G. A., Young, R. S., Reichenauer, K., Stevens, J., and Ventrone, F. (2004). Gps multipath
mitigation assessment of digital beam forming antenna technology in a jpals dual frequency
smoothing architecture. In Proceedings of the 2004 National Technical Meeting of the Institute of
Navigation, pp. 561–572.
27. Park, K. D., Nerem, R. S., Schenewerk, M. S., & Davis, J. L. (2004). Site-specic multipath
characteristics of global IGS and CORS GPS sites. Journal of Geodesy, 77(12), 799-803.
https://doi.org/10.1007/s00190-003-0359-9
2. Rebischung, P., Altamimi, Z., Ray, J., & Garayt, B. (2016). The IGS contribution to ITRF2014. Journal
of Geodesy, 90(7), 611-630. https://doi.org/10.1007/s00190-016-0897-6
Page 17/28
29. Satirapod, C. and Rizos, C. (2005). Multipath mitigation by wavelet analysis for GPS base station
applications. Survey Review, 38(295):2–10. https://doi.org/10.1179/003962605791521699
30. Su, M., Yang, Y., Qiao, L., Teng, X., & Song, H. (2021). Enhanced multipath mitigation method based
on multi-resolution CNR model and adaptive statistical test strategy for real-time kinematic PPP.
Advances in Space Research, 67(2), 868-882. https://doi.org/10.1016/j.asr.2020.10.035
31. Suzuki, T., Kusama, K., & Amano, Y. (2020). NLOS Multipath Detection using Convolutional Neural
Network. In Proceedings of the 33rd International Technical Meeting of the Satellite Division of The
Institute of Navigation (ION GNSS+ 2020), pp. 2989-3000. https://doi.org/10.33012/2020.17663
32. Takasu, T. (2009). RTKLIB: Open source program package for RTK-GPS. Proceedings of the FOSS4G.
33. Tao, Y., Liu, C., Chen, T., Zhao, X., Liu, C., Hu, H., Zhou, T., Xin, H., & Neagu, A. (2021). Real-Time
Multipath Mitigation in Multi-GNSS Short Baseline Positioning via CNN-LSTM Method. Mathematical
Problems in Engineering, 2021, 1-12. https://doi.org/10.1155/2021/6573230
34. Tao, Y., Liu, C., Liu, C., Zhao, X., Hu, H., & Xin, H. (2021). Joint time–frequency mask and
convolutional neural network for real-time separation of multipath in GNSS deformation monitoring.
GPS Solutions, 25(1). https://doi.org/10.1007/s10291-020-01074-y
35. Van Dierendonck, A., Fenton, P., and Ford, T. (1992). Theory and performance of narrow correlator
spacing in a GPS receiver. Navigation, 39(3):265–283.
3. Wang, Z., Chen, W., Dong, D., Wang, M., Cai, M., Yu, C., Zheng, Z., & Liu, M. (2019). Multipath
mitigation based on trend surface analysis applied to dual-antenna receiver with common clock.
GPS Solutions, 23(4). https://doi.org/10.1007/s10291-019-0897-0
37. Xu, H., Angrisano, A., Gaglione, S., & Hsu, L.-T. (2020). Machine learning based LOS/NLOS classier
and robust estimator for GNSS shadow matching. Satellite Navigation, 1(1).
https://doi.org/10.1186/s43020-020-00016-w
3. Ye, S., Chen, D., Liu, Y., Jiang, P., Tang, W., & Xia, P. (2014). Carrier phase multipath mitigation for
BeiDou navigation satellite system. GPS Solutions, 19(4), 545-557. https://doi.org/10.1007/s10291-
014-0409-1
39. Zheng, K., Zhang, X., Li, P., Li, X., Ge, M., Guo, F., Sang, J., & Schuh, H. (2019). Multipath extraction and
mitigation for high-rate multi-GNSS precise point positioning. Journal of Geodesy, 93(10), 2037-
2051. https://doi.org/10.1007/s00190-019-01300-7
40. Zhong, P., Ding, X., Yuan, L., Xu, Y., Kwok, K., & Chen, Y. (2009). Sidereal ltering based on single
differences for mitigating GPS multipath effects on short baselines. Journal of Geodesy, 84(2), 145-
158. https://doi.org/10.1007/s00190-009-0352-z
Figures
Page 18/28
Figure 1
Flow chart of GPS data processing and ML-based multipath mitigation. The ML model is trained using
the data from previous days and it predicts the multipath errors for the target day
Page 19/28
Figure 2
(a) GPS station distribution in bird’s-eye view. CUT0 is the reference station used for relative positioning.
All three rover antennas are connected to two different receivers, respectively. (b) View of the observation
environment from the rooftop of Curtin University. The copyright of the photos are preserved by Curtin
GNSS-SPAN Group
Page 20/28
Figure 3
Low-pass ltered residuals of satellite G04 at station CUCC. Pseudorange and carrier phase residuals are
shown in the upper and bottom panels, respectively. Residuals on DOY 244, 245 and 249 are denoted by
red, blue and green curves, respectively, and they are shifted by corresponding orbit repeat periods along
the x-axis to better show the temporal correlation. Shifts along the y-axis are made to avoid overlapping.
Pearson correlation coecients with respect to the residuals of DOY 244 are denoted above each curve
for DOY 245 and 249
Page 21/28
Figure 4
Mean residual reduction rates for station CUCC over DOY 250-273. The reduction rates using MLP, RF and
XGB (only azimuth and elevation as input features) are denoted by blue, orange and green bars,
respectively. The shaded bars represent the corresponding models with SNR as an additional input
feature
Page 22/28
Figure 5
Visualization of ML-based (XGB) multipath models for station CUCC on DOY 251
Page 23/28
Figure 6
(a) Low-pass lterd residuals and different multipath models for satellite G10 at station CUCC on DOY
251. The ltered residuals and multipath models of SF, MHM and XGB are displayed in black, red, blue
and green, respectively. The individual curves are shifted vertically to avoid overlapping. (b) L1 power
spectral density (relative to 1 m2/Hz) for low-pass ltered residuals and different multipath models. The
two orange vertical lines represent the frequency range from 17 s to 60 s
Page 24/28
Figure 7
Raw and multipath corrected residuals of satellite G10 at station CUCC on DOY 251. The raw residuals
and those corrected with SF, MHM and XGB are represented by black, red, blue and green curves,
respectively. The individual curves are shifted along the y-axis to avoid overlapping. The RMS value is
denoted above each curve
Page 25/28
Figure 8
Mean residual reduction rates using different multipath mitigation methods over all ve stations and 25
days
Page 26/28
Figure 9
Comparison of validity periods for different multipath models. Data from DOY 244-248 is used to set up
the multipath models, which are later applied for multipath mitigation on DOY 249-273. The residual
reduction rate on each day is the mean value of all ve stations
Page 27/28
Figure 10
Hz kinematic positioning results at station CUCC on DOY 273 with multipath mitigated with different
methods. The raw positioning result and those corrected with SF, MHM and XGB are shown in black, red,
blue and green curves, respectively. RMS values of displacements are denoted above each curve. Only the
last 6 h displacements are displayed to show the detailed positioning errors induced by multipath
Page 28/28
Figure 11
Residual reduction rates for 30 s sampling data using different multipath mitigation methods. The
residual reduction rate on each day is the mean value of all ve stations
... While from the positional perspective, modeling accuracy is evaluated through positioning performance. For instance, when a machine learning technique is used to model multipath faults (Pan et al. 2023), any improvement in position accuracy is considered an indicator of the accuracy of GNSS fault modeling. However, there are several shortcomings in the current techniques for GNSS fault model evaluation. ...
Article
Full-text available
The characteristics of residual errors in GNSS positioning are crucial for fault detection and integrity monitoring. Despite the wide use of the zero-mean Gaussian assumption in the navigation community, studies highlight non-Gaussian traits and heavy-tailed patterns in residual errors. The problem will be even more challenging for users in difficult environments where residual errors consist of a combination of multiple modes with high complexity and cannot be fitted with known distributions or empirical models. To address these issues, our work introduces a novel approach leveraging the Wasserstein distance for assessing the performance of error characterization and fault modeling. However, relying solely on the Wasserstein distance value for direct similarity assessment is hindered by its dependency on dimensionality. We propose a second-order Gaussian Wasserstein distance-based precision metric to offer a quantitative evaluation of GNSS error models in terms of both goodness-of-fit and underlying assumptions. We also establish a robust scoring criterion to distinguish between various GNSS error models, ensuring comprehensive evaluation. The proposed method is validated through a known high-dimensional Gaussian model, achieving a score of 99.95 over 100 with a sample size of 10,000. To demonstrate the capability in dealing with complexity, two multivariate complex GNSS models incorporating copula functions to capture intricate inter-dimensional correlations are established and assessed by our approach. Experimental results show that the method can effectively deliver the evaluation of goodness-of-fault models using the establishment of a universal criteria with different dimensions. It provides a quantitative measure on the goodness of fittings and enhances the modeling to reflect the reality, therefore solving the problems raised above. In addition, with this technique, the close-to-reality fault models can be chosen to generate simulated faulty datasets, thus benefiting algorithm testing and improvement. This is also beneficial to more accurate integrity risk assessment to avoid overbounding- or underbounding-resulted false or missed alert.
Article
Full-text available
Remote sensing of water vapour using the Global Navigation Satellite System (GNSS) is a well-established technique and reliable data source for numerical weather prediction (NWP). However, one of the phenomena rarely studied using GNSS are foehn winds. Since foehn winds are associated with significant humidity gradients between two sides of a mountain range, tropospheric estimates from GNSS are also affected by their occurrence. Time series reveal characteristic features like distinctive minima and maxima as well as a significant decrease in the correlation between the stations. However, detecting such signals becomes increasingly difficult for large datasets. Therefore, we suggest the application of machine learning algorithms for the detection and prediction of foehn events by means of GNSS troposphere products. This initial study develops a new, machine learning-based method for detection and prediction of foehn events at the Swiss station Altdorf by utilising long-term time series of high-quality GNSS troposphere products. Data from the Automated GNSS Network Switzerland (AGNES) and various GNSS sites from neighbouring countries as well as records of an operational foehn index are used to investigate the performance of several different classification algorithms based on appropriate statistical metrics. The two best-performing algorithms are fine tuned and tested in four dedicated experiments using different feature setups. The results are promising, especially when reprocessed GNSS products are utilised and the most dense station setup is used. Detection- and alarm-based measures reach levels between 66 %–80 % for both tested algorithms and thus are comparable to those from studies using data from meteorological stations and NWP. For operational prediction, limitations due to the availability and quality of GNSS products in near-real time (NRT) exist. However, they might be mitigated to a significant extent by provision of additional NRT products and improved data processing in the future. Results also outline benefits for the results when including geographically relevant stations (e.g. high-altitude stations) in the utilised datasets.
Article
Full-text available
Machine learning and, more precisely, data-driven models are providing solutions where physics-based models are intractable. This article discusses the use of deep learning models to characterize the intricate effects of multipath propagation on GNSS correlation outputs. Particularly, we aim at substituting standard correlation schemes, optimal under single-ray Gaussian noise assumptions, with Neural Network (NN)-based correlation schemes, that are able to learn the otherwise challenging to model multipath channels. The paper shows that Deep Neural Networks (DNNs), as applied to tracking loops, can provide enhanced performance as compared to standard correlation schemes in $i)$ Line-of-Sight (LOS) scenarios, by filtering out more noise thanks to strong prior regularization through knowledge of correlation characteristics and Gaussian noise during training process; and $ii)$ at the same time, the DNN can adjust its behavior to better disentangle multipath signals from LOS signals. This article provides results showing the superiority of the proposed DNN trained models, with focus on time-delay tracking in a variety of realistic scenarios.
Article
Full-text available
Global navigation satellite systems (GNSS) provide globally distributed station coordinate time series that can be used for a variety of applications such as the definition of a terrestrial reference frame. A reliable estimation of the coordinate time series trends gives valuable information about station movements during the measured time period. Detecting discontinuities of various origins in such time series is crucial for accurate and robust velocity estimation. At present, there is no fully automated standard method for detecting discontinuities. Instead, discontinuity-catalogues are frequently used, which provide information about when a device was changed or an earthquake occurred. However, it is known that these catalogues suffer from incompleteness. This study investigates the suitability of machine learning classification algorithms that are fully data-driven to detect discontinuities caused by earthquakes in station coordinate time series without the need for external information. For this study, Japan was selected as a testing area. Ten different machine learning algorithms have been tested. It is found that Random Forest achieves the best performance with an F1 score of 0.77, a recall of 0.78, and a precision of 0.76. Overall, 525 of 565 recorded earthquakes in the test data were correctly classified. It is further highlighted that splitting the time series into chunks of 21 days leads to the best performance. Furthermore, it is beneficial to combine the three (normalized) components of the GNSS solution into one sample, and that adding the value range as an additional feature improves the result. Thus, this work demonstrates how it is possible to use machine learning algorithms to detect discontinuities in GNSS time series.
Article
Full-text available
A new generation of earthquake catalogs developed through supervised machine-learning illuminates earthquake activity with unprecedented detail. Application of unsupervised machine learning to analyze the more complete expression of seismicity in these catalogs may be the fastest route to improving earthquake forecasting.
Article
Full-text available
The multipath effect reduces the accuracy of pseudorange and carrier phase observations and significantly affects the convergence of precise point positioning (PPP) in an urban high-obstruction environment. The multipath hemispherical map (MHM) is based on the spatial repeatability of the multipath and is used to mitigate the multipath effect. This method is widely used because it is suitable for real-time dynamic and static situations with multipath invariance. Although the MHM algorithm is simple and easy to implement, it ignores the spatial distribution of multipath inside the sky grid. It is not suitable for use in high-frequency multipath corrections. By fitting the multipath trend inside the sky grid, the MHM based on a trend-surface analysis (T-MHM) alleviates the high-frequency and low-frequency multipath simultaneously, improving the accuracy of the baseline solution in the differential mode. We mainly demonstrate the application of T-MHM in PPP, analyze the unmodeled errors, evaluate the multipath correction effect of T-MHM, discuss the optimal modeling days, and test its sensitivity in the scale of the sky grid. Compared to MHM, the experimental results show that using T-MHM in multipath correction improves the positioning accuracy in the horizontal and vertical directions by 21.6 and 13.6%, respectively, and shortens the convergence time by 24.2 and 7.2%, respectively. The T-MHM method is not sensitive to the grid scale of the sky, thus resulting in high efficiency.
Article
Full-text available
This study proposes an enhanced multipath mitigation method based on multi-resolution carrier-to-noise-ratio (CNR) model and adaptive statistical test strategy for real-time kinematic precise point positioning (PPP) applications. The multi-resolution CNR model is established with GPS observation data collected from DOY 152 to 181 of 2019 by 230 globally distributed IGS stations, which used to analyze the relevant factors affecting CNR. Statistical results indicate that the CNR is not only related to the satellite elevation, but also closely related to the receiver types and specific satellite. The maximum difference between different receivers can reach 20 dB for the same satellite at the same elevation. In addition, the performance of the CNR is also obviously different between each satellite, and the maximum difference between different satellites is about 10 dB for the same receiver at the same elevation. Hence, in terms of the method which is based on CNR information for multipath detection and mitigation, the independence of receiver types, satellite and frequency must be considered. With the above analysis, this study developed a multi-resolution CNR model based on different receiver types, different satellites and different elevation firstly. Then, combined with the adaptive statistical test strategy which is based on the difference of CNR between inter-frequency and the difference of CNR between adjacent epochs, the multipath can be detected effectively. For the epoch which affected by multipath, the down-weighted strategy based on CNR is adopted to mitigate the influence of multipath on positioning. Real-time kinematic PPP data are collected to assess the proposed method, and the results demonstrate that the proposed method can detect the multipath effectively, and the detection rate can reach 90.28%. Moreover, after adopting the mitigation strategy, the RMS bias of the east, north and up components are improved about 19.95%, 17.89% and 23.07% compared to the original results, respectively. It is worth noting that this proposed method is also suitable for other GNSS, such as GLONASS and BDS, but the corresponding CNR model must be established simultaneously.
Article
Full-text available
Global Navigation Satellites Systems (GNSS) is frequently used for positioning services in various applications, e.g., pedestrian and vehicular navigation. However, it is well-known that GNSS positioning performs unreliably in urban environments. GNSS shadow matching is a method of improving accuracy in the cross-street direction. Initial position and classification of observed satellite visibility between line-of-sight (LOS) and non-line-of-sight (NLOS) are essential for its performance. For the conventional LOS/NLOS classification, the classifiers are based on a single feature, extracted from raw GNSS measurements, such as signal noise ratio (SNR), pseudorange, elevation angle, etc. Especially in urban canyons, these measurements are unstable and unreliable due to the signal reflection and refraction from the surrounding buildings. Besides, the conventional least square (LS) approach for positioning is insufficient to provide accurate initialization for shadow matching in urban areas. In our study, shadow matching is improved using the initial position from robust estimator and the satellite visibility determined by support vector machine (SVM). The robust estimator has an improved positioning accuracy and the classification rate of SVM classification can reach 91.5% in urban scenarios. An important issue is related to satellites with ultra-high or low elevation angles and satellites near the building boundary that are very likely to be misclassified. By solving this problem, the SVM classification shows the potential of about 90% classification accuracy for various urban cases. With the help of these approaches, the shadow matching has a mean error of 10.27m with 1.44m in the cross-street direction; these performances are suitable for urban positioning.
Article
Full-text available
Multipath effect on carrier-phase observation is one of the bottlenecks for mm-level applications when using precise point positioning (PPP). Hence, we extract the multipath directly from raw carrier-phase residuals of GPS, GLONASS, Galileo, and BDS, by using PPP technique. Although the residuals for one frequency assimilate the errors from other frequencies, which is caused by error adjustment by the least squares estimator, the primary component of residuals is multipath. The results indicate that the residuals between frequencies have a significant linear negative correlation and synchronous time lag for each system. Besides BDS Geostationary Earth Orbit satellites, the residuals for other satellites can establish accurate mathematic relationship between the frequencies. For GLONASS, the residuals of R1 frequency recovered from R2 frequency with the mathematical relationship are better than 0.1 mm accuracy, which means the effect of inter-frequency bias can be neglected. These regularities double-reduce the complexity of data processing. Based on the multipath distribution, we propose a modified Multipath Hemispherical Map model (M-MHM), which constructs grids from residuals and is divided into three equal-elevation angle parts with an optimal resolution 0.2° × 0.2° × 1° from numerous experiments. In addition, the multipath manifests great consistency among satellites for GPS, GLONASS, and Galileo systems when elevation angles are higher than 15°, while is more satellite dependent for BDS. Although GPS L1 frequency is identical to Galileo E1, the model still has some systematic bias between GPS and Galileo. Compared with sidereal filtering and original MHM model, the M-MHM brings the highest improvement in both residual variance reduction and positioning accuracy. The positioning accuracy is on average 12% improvement compared to MHM and 29% improvement compared to SF. For four systems combined solutions with the M-MHM model, can reach an accuracy of 0.75, 0.55, and 2.08 cm in the east, north, and up components.
Article
Full-text available
Multipath mitigation methods based on the intrinsic spatiotemporal repeatability of multipath effect are commonly applied in high-precision Global Navigation Satellite System positioning. Among which multipath hemispherical map (MHM) has the advantages of simplicity of calculation and real-time multipath reduction of static environment or some dynamic environment. However, utilizing the average of residuals as the calibration values will inevitably lead to insufficient or excessive modification of multipath in the lattice. We proposed an advanced multipath elimination method named trend surface analysis-based multipath hemispherical map (T-MHM), which extends the primary MHM by combining the trend surface analysis method to investigate the multipath spatial distribution and trend in each lattice. A short baseline experiment was carried out using a dual-antenna receiver with common clock in campus of East China Normal University. Single-difference carrier residuals were used for relative multipath modeling. The multipath elimination results indicate that, compared with the original approach, the T-MHM improves the dispersion and accuracy of the baseline solution and increases the phase observation residuals reduction rate. It also enhances the modeling of multipath information in both low- and high-frequency bands. Meanwhile, T-MHM is less sensitive to lattice size. In practical applications, larger lattice size can be adopted to improve the computational efficiency and spatial coverage with little loss of accuracy.