Next Article in Journal
Seasonal Variation of Dust Aerosol Vertical Distribution in Arctic Based on Polarized Micropulse Lidar Measurement
Next Article in Special Issue
The Temporal Evolution of F-Region Equatorial Ionization Anomaly Owing to the 2022 Tonga Volcanic Eruption
Previous Article in Journal
The Time Lag Effect Improves Prediction of the Effects of Climate Change on Vegetation Growth in Southwest China
Previous Article in Special Issue
Analysis of the Ionospheric Irregularities and Phase Scintillation at Low and Middle Latitudes Based on Swarm Observations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Machine Learning-Based Method for Modeling TEC Regional Temporal-Spatial Map

1
School of Microelectronics, Tianjin University, Tianjin 300072, China
2
Qingdao Institute for Ocean Technology, Tianjin University, Qingdao 266200, China
3
Shandong Engineering Technology Research Center of Ocean Information Awareness and Transmission, Qingdao 266200, China
4
College Electronic and Information, Qingdao University, Qingdao 266071, China
5
Tianjin Key Laboratory of Imaging and Sensing Microelectronic Technology, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Remote Sens. 2022, 14(21), 5579; https://doi.org/10.3390/rs14215579
Submission received: 4 October 2022 / Revised: 29 October 2022 / Accepted: 2 November 2022 / Published: 4 November 2022

Abstract

:
In order to achieve the high-accuracy prediction of the total electron content (TEC) of the regional ionosphere for supporting the application of satellite navigation, positioning, measurement, and controlling, we proposed a modeling method based on machine learning (ML) and use this method to establish an empirical prediction model of TEC for parts of Europe. The model has three main characteristics: (1) The principal component analysis (PCA) is used to separate TEC’s temporal and spatial variation characteristics and to establish its corresponding map, (2) the solar activity parameters of the 12-month mean flux of the solar radio waves at 10.7 cm (F10.712) and the 12-month mean sunspot number (R12) are introduced into the temporal map as independent variables to reflect the temporal variation characteristics of TEC, and (3) The modified Kriging spatial interpolation method is used to achieve the spatial reconstruction of TEC. Finally, the regression learning method is used to determine the coefficients and harmonic numbers of the model by using the root mean square error (RMSE) and its relative value (RRMSE) as the evaluation standard. Specially, the modeling process is easy to understand, and the determined model parameters are interpretable. The statistical results show that the monthly mean values of TEC predicted by the proposed model in this paper are highly consistent with the observed values curve of TEC, and the RRMSE of the predicted results is 12.76%. Furthermore, comparing the proposed model with the IRI model, it can be found that the prediction accuracy of TEC by the proposed model is much higher than that of the IRI model either with CCIR or URSI coefficients, and the improvement is 38.63% and 35.79%, respectively.

Graphical Abstract

1. Introduction

The ionosphere is the part of the upper atmosphere between 60 km to 1000 km above the ground. It is affected by many factors, such as solar radiation, geomagnetic disturbance, and atmospheric activity. Furthermore, with the change in geographical location and solar activity, the ionosphere presents the inherent periodic variation of day, month, season, year, and complex synoptic and climatological changes [1]. In general, when observational radio waves transmitted by space geodetic techniques such as global navigation satellite system (GNSS) [2] and very long baseline interferometry (VLBI) [3] pass through the ionosphere, their phase velocity and group velocity due to the collision of free electrons in the ionosphere will produce a certain degree of interference, which is approximately proportional to the total electron content (TEC) along the signal propagation path [4]. The resulting delay effect can be up to tens of meters [5], thus mastering TEC is very crucial for these systems such as GNSS and VLBI, etc. Supposing we can master the distribution of electron content in the ionosphere, we will not only be quantitatively able to study the influence of the ionosphere on the propagation of radio wave signals but also able to summarize the variation laws of the ionosphere at different time and space scales. Furthermore, we will achieve advanced prediction and real-time processing of ionospheric disturbances and reduce the attenuation and variation in the signal propagation process.
Considering the importance of accurately reflecting the distribution of electron content in the ionosphere for improving applications such as satellite communication, navigation, positioning [6], measurement, and disaster prevention [7,8], TEC observation, research, and modeling have always been a hot topic of the exploration for decades. TEC can be directly measured by the ionospheric sounder or indirectly inverted by carrier phase delay during the transmission of the radio signal [9]. In the absence of measurement conditions or other extreme cases, TEC can be calculated by physical models [10,11], mathematical models [12,13], and empirical models [14,15]. Among them, the international reference ionosphere (IRI) is the most common empirical model and is also a well-known recommended international standard, which is often used as a comparison model to verify other models [16,17]. The IRI model has been steadily improved with newer data and better modeling techniques by the joint working group of the Committee on Space Research and the International Union of Radio Science (URSI) [18]. For some ionospheric parameters such as TEC and the critical frequency of the F2 layer (foF2), there are two sub-models with the corresponding coefficients provided by the URSI and the Consultative Committee of International Radio (CCIR). The latest version is IRI-2016 [18], which determines the electron density by extending the E and F-layer’s anchor points of peak electron densities. In particular, the maximum electron density is calculated by the NeQuick model [19].
The recent research has the following properties in modeling TEC. On the one hand, machine learning (ML) and deep learning (DL) methods have been introduced into the modeling of ionospheric parameters and played a good role in improving the accuracy of empirical models [20]. For example, Kaselimi et al. [21] propose a recurrent neural network (RNN) model based on deep learning to predict TEC. Some ionospheric TEC prediction models based on long short-term memory (LSTM) were proposed to predict TEC [22,23], and some ionospheric TEC maps have been proposed based on bidirectional LSTM algorithms in different regions [24,25]. A TEC prediction model based on a multi-step auxiliary LSTM algorithm was established and can effectively alleviate the increasing error with prediction time [26]. The TEC forecasting methods based on generative adversarial networks (GAN) [27,28] and conditional GAN [29] were proposed and improved the IRI [30]. At the same time, Han et al. have modeled ionospheric TEC by four different ML methods such as artificial neural network (ANN), LSTM networks, adaptive neuro-fuzzy inference system, and gradient boosting decision tree (GBDT) [31]. In addition, the nearest neighbor method [32], prophet model [33], and decision tree [34] based on ML are also introduced into the modeling of ionospheric TEC, providing new modeling ideas and excellent prediction results. On the other hand, compared with the global model, the regional model has made great progress and provided higher prediction accuracy. Therefore, to achieve higher accuracy prediction of specific regions, TEC prediction models focusing on local regions have gradually become a research hotspot in recent years. Recent achievements have finished in Japan [35], China [36], Korean Peninsula [37], African Region [38], Antarctic Region [39], the low latitude region [2], the sea region [31], etc. For example, a European TEC prediction model [40] based on thin-plate splines (TPS) interpolation was self-calibrated, making the model better accurate during geomagnetic storms. El-Diasty [41] combined a neural network (NN) algorithm and wavelet decomposition algorithm to establish a short-term forecast model of TEC, which can adapt to various ionospheric environments more flexibly and has good robustness. A regional TEC model was established using the single-layer model and proved superior to the IRI model [42]. Constantly updated data measurement methods, data processing techniques, and modeling methods are designed to improve TEC prediction accuracy for military and civilian applications.
The TEC estimating methods can divide into the Long-term Prediction Method (LTPM) and the Short-term Forecast Method (STFM). LTPM is often used to predict the monthly statistics of VTEC, and STFM refers to forecasting TEC for several hours or days in the future. Unlike STFM, which pays more attention to the diurnal variation of TEC, LTPM pays more attention to the periodic variation trend of TEC monthly median value [43]. Considering the previous investigation, the prediction accuracy of the IRI model in some parts of Europe was low due to the complex dynamic change characteristics of the ionosphere in Europe, which needed to be urgently improved [42]. Compared with previously used ML methods such as RNN [21], LSTM networks [26], traditional NN [35], etc., the proposed ML-based modeling method is to make the model interpretable and simple. To improve the TEC long-term prediction accuracy in Europe, we propose an ML-based modeling method and use this method to establish an empirical prediction model of TEC for parts of Europe. This paper is organized as follows. First, Section 2 will specifically describe model creation and parameters determination. Then, the validation of the model and comparison with other models will be presented in Section 3. Finally, Section 4 will summarize the article and propose the direction for further improvement in the follow-up work.

2. Modeling

2.1. Modeling Method

Modeling methods of the TEC model can be divided into traditional and intelligent methods. Most traditional methods are based on mathematical principles; that is, different mathematical functions or spatial interpolation methods are used to characterize the temporal-spatial dynamic variations of TEC. However, with the rapid development of artificial intelligence, many intelligent methods are beginning to emerge, the methods based on ML are introduced into the model of TEC and organic combination with traditional methods to promote the leapfrog development of the TEC model. ML is interdisciplinary and builds a probabilistic statistical model based on data and uses the model to predict and analyze data [44]. Compared with the black box algorithm of the NN [45], the data processing ideas of ML are easy to understand, and the determined model parameters can be interpreted [46]. Therefore, ML is widely applied in the modeling of various ionospheric parameters. As shown in Figure 1, ML is a process of determining the three elements (model, algorithm, and strategy) and solving the following four problems:
  • What kind of data are needed for training and predicting? The premise of ML is that the data used for training models have the same statistical characteristics, and the modeling process is to find the appropriate map to describe such statistical characteristics. This paper selects the monthly mean values of TEC (hereinafter referred to as TEC) from eight ionospheric observation stations in parts of Europe as the modeling dataset. The validation dataset consists of six stations in the same region, of which three are used for modeling, and three are not involved in the modeling process. The details are described in Section 2.2.
  • How to choose a suitable model? This requires us to have an in-depth understanding of the purpose of modeling, the principles and applicable scope of different models, their advantages and disadvantages, and the characteristics of the training dataset before modeling. In this paper, the principal component analysis (PCA) function set is selected as the hypothesis space to construct the temporal and spatial characteristic map of TEC. PCA is a commonly used data analysis method known as eigenvector analysis [47]. It transforms the original data into a set of linearly independent ordered basis functions and correlation coefficients through a linear transformation. It can extract the main eigen components of the data and is often used for dimensionality reduction in high-dimensional data.
In 1901, Person [48] first proposed PCA as an effective method to establish an empirical model. And then, Dvinskikh [49] introduced it into the modeling of the ionospheric parameter empirical model for the first time in 1988. Subsequently, many empirical models of ionospheric parameters, such as foF2 [50], the propagation factor at 3000 km of the F2 layer [51], and TEC [52], were modeled using the PCA method. PCA can not only reduce the number of parameters involved in the modeling process but also reduce the amount of calculation and storage of intermediate data to achieve high accuracy and fast convergence [53]. Therefore, we apply PCA to decompose TEC orthogonally into temporal and spatial variables controlled by independent and unrelated processes so that TEC can be calculated through the orthogonal temporal and spatial map. The flow chart of modeling and validation is shown in Figure 2.
The nonlinear map of geographical location (latitude and longitude), solar activity, month, and universal time to TEC (unit is TECU) constructed based on PCA theory can be expressed as the harmonic function:
TEC ( λ , φ , p , r , m , t ) = n = 1 N W n ( λ , φ ) T ^ EC ( p , r , m , t ) ,
where λ is the geographical latitude, φ is the geographical longitude, p is the average solar radio flux of 10.7 cm in 12 months (F10.712), r is the average number of sunspots in 12 months (R12), m is the month, t is the given universal time, N is the number of ionospheric observation stations used for modeling, T ^ EC ( p , r , m , t ) is the temporal map of TEC, and Wn(λ, φ) is the weight coefficients of the spatial map of TEC (details are provided in Section 2.3).
  • How to determine the parameters of the model? ML is a supervised learning method [44] that obtains the target model and the corresponding algorithm through training. The least-squares (LS) regression analysis algorithm is one of the criteria used to determine the parameters of the regression model. Due to its advantages of simple calculation, short modeling time, easy understanding, and being highly sensitive to outliers, this algorithm has been widely used in parameter prediction and time series modeling in many fields. And LS shows good performance; therefore, this is also the reason why we choose it for modeling.
  • How to evaluate the performance of the model? First, we must develop a unified standard to determine the optimal model and its correlation coefficients from the hypothesis space. Therefore, we calculate the following parameters as the evaluation standard:
    (a)
    Root mean square error (RMSE):
    RMSE = 1 H h = 1 H ( TEC o h TEC p h ) 2 ,
    (b)
    Relative root mean square error (RRMSE):
RRMSE = 1 H h = 1 H ( TEC o h TEC p h TEC o h ) 2   × 100 ,
where TECoh is the hth observed value of TEC, TECph is the hth predicted value of TEC, and H is the total statistical amount.

2.2. Modeling Dataset

As previously mentioned, ML is a data-driven modeling approach. In simple terms, the process of ML can be summarized as determining the learning model, strategy, and algorithm based on the determined dataset, finally finding the optimal model through training, then using the proposed model to predict the unknown data. Therefore, the selection of appropriate modeling data is an important prerequisite to ensuring the accuracy of modeling.
The data used for modeling and validation in this paper are obtained from the National Oceanic and Atmospheric Administration (NOAA) and the Global Ionospheric Radio Observatory (GIRO). This paper focuses on exploring an ML-based method for modeling the TEC temporal-spatial map, and the TEC provided by NOAA and GIRO is input data to test this method’s performance. These data have good credibility and were selected for modeling, as reported by [54]. The data information is listed in Table 1. We can download the required data by selecting the corresponding time, observed station, and data type, through the official websites of NOAA [55] and GIRO [56]. The downloaded data can be saved as a text file in which recorded 24 h of TEC data for each day of the year at a station; that is, each line gives the TEC at one moment. Specifically, part of the data of the Pruhonice, Roquetes, Rome, etc., stations are obtained through NOAA, part of the data of the Athens, Fairford, Dourbes, etc., stations, and all the data of the Sopron and Nicosia stations used for validation are obtained through GIRO.
Figure 3 shows the geographical distribution of the 11 ionospheric stations used for modeling and validation in this paper in parts of Europe. The monthly mean valid data volume of TEC is 42,049 in total. To ensure the data continuity and the model’s validity, we selected that the continuous period of each station’s valid data is longer than one solar activity cycle, which covers at least one low solar activity year and one high solar activity year (as shown in Figure 4). As mentioned above, the observation data of different periods at the same station are obtained from NOAA and GIRO, respectively. Considering that the TEC data provided by GIRO has a longer continuous period and a larger amount, the missing periods in GIRO were supplemented by NOAA and then merged into a group for modeling and validation. For some stations whose valid data continuous period is less than one solar activity cycle, such as the Sopron, Nicosia, and Chilton, we take them as validation stations and not as modeling stations to participate in the modeling process. For the collected data, we first preprocessed them in the following two steps:
(1)
Considering that the observation period of different stations includes 15 min, 30 min, and 60 min, to ensure the uniformity of data format and avoid introducing errors, we choose 60 min as the sampling period to ensure data consistency.
(2)
In order to meet the modeling requirements of the Long-term regional ionospheric TEC prediction model, the monthly 24-h TEC mean values of each TEC observation station are calculated, and one observation station corresponds to a standardized modeling data file. After the above processing, we selected 8 stations for modeling and 6 stations for validation. Among the validation stations, three participated in modeling and three did not.

2.3. Modeling Determination

2.3.1. Modeling of Temporal Characteristics

It is well known that photoionization is the main mechanism of ionospheric plasma generation, and solar radiation’s energy mainly determines photoionization’s intensity. The main parameters used to describe solar activity are the flux of the solar radio waves at 10.7 cm (F10.7) and sunspot number (R), which show similar periodic variation due to the influence of the corona and chromosphere. The 12-month mean F10.7 (F10.712) is used to characterize the energy of extreme ultraviolet radiation (EUV) in the empirical model of solar radiation intensity [59]. The 12-month mean R (R12) reflects the average intensity of solar activity. Specifically, in a month, F10.712 has 24 data points, corresponding to the 24 h of a day; that is, each hour corresponds to a value of F10.712. The R12 and TEC are consistent with F10.712. Studies have shown that F10.712 and R12 can better reflect photoionization intensity than daily F10.7 and R [60]. In addition, Sivavaraprasad [61] also used F10.712 and R12 to establish a TEC prediction model suitable for low latitude areas of the equator, which was proved to have a great prediction effect, further illustrating the correctness and applicability of using F10.712 and R12 to establish a long-term TEC prediction model.
Figure 5 compares F10.712, R12, and TEC in the same period using the data of the Fairford station at UT 0:00 from 2004 to 2020. It is evident that TEC keeps the same variation cycle with F10.712 and R12; that is, rising or falling at the same time points. However, the variation of TEC in detail is slightly different from that of F10.712 and R12. The value of TEC increased significantly in high solar activity years, showing prominent annual variation characteristics, which are also drawn in [62,63]. Furthermore, we calculated the correlation coefficients between TEC and different solar activity parameters in 12 months and 24-h, which are given in Figure 6. The circle size corresponds to the correlation coefficients value. It can be seen that the corresponding correlation coefficients of the three solar activity parameters and TEC are all concentrated above 0.8, indicating that the three solar activity parameters have an obvious correlation with TEC. However, there are still some differences in the correlation coefficients of different parameters at the same time points. For example, at 19h in February, the correlation coefficient between F10.712 and TEC is significantly greater than that of R12.
Considering the coupling effect of two solar activity parameters on TEC in different aspects, F10.712 and R12 are introduced into the temporal model together as independent variables. This paper focuses on the prediction and analysis of TEC, so we assume that the prediction of solar activity parameters is known [64,65]. The time resolution of the temporal model is 1h. Given specific geographic coordinates and universal time, we can establish a temporal map from solar activity parameters and periodic oscillations to TEC, as shown in (4):
TEC = T ^ EC ( p , r , m , t )                   = k = 0 K l = 0 L [ c k , l p p l cos ( 2 π k m / 12 ) + s k , l p p l sin ( 2 π k m / 12 ) + ,                     c k , l r r l cos ( 2 π k m / 12 ) + s k , l r r l sin ( 2 π k m / 12 ) ]
where p is F10.712, r is R12, m is the month, t is the given universal time, K represents the upper bound of the time cycle, with values of 1, 2, 3, and 4 respectively representing annual, semi-annual, seasonal, and monthly cycle. L represents the periodic variation of the solar activity; that is, the l-order harmonic of the solar activity index is taken for regression. The values of K and L are determined by regression analysis. In addition, the coefficients c k , l p , s k , l p , c k , l r , s k , l r are determined by the LS regression.

2.3.2. Modeling of Spatial Characteristics

TEC shows the irregular and uneven distribution in space, and the existing ionospheric observation methods cannot achieve global observation. Therefore, we can obtain several observed samples from ionospheric observation stations in different regions. Furthermore, these observations can reflect part of the characteristics of ionospheric TEC spatial distribution and can predict the characteristics of TEC in the unknown regions accordingly.
Considering the modeling data volume, the spatial distribution of observations, and the final purpose of modeling, we selected the modified Kriging interpolation method as the basic algorithm for spatial characteristics modeling. And then, it is further optimized according to the modeling data to obtain a better prediction result. The specific process is shown in Figure 7.
Firstly, the covariance function is generated by using the regionalization variables that satisfy second-order stationary and eigen hypothesis. Then, we can determine the variation function using the ionospheric distance between the observation and reconstruction stations. Namely, we can estimate the spatial autocorrelation between TEC samples. Secondly, with the minimum variance as the standard, the weight coefficients of different observation stations are solved to determine the Kriging interpolation function. Finally, the determined Kriging interpolation function is used to calculate the TEC of the reconstruction stations, and the continuous TEC data plane is generated.
In summary, the general expression of TEC spatial reconstruction based on modified Kriging [66] can be written as follows:
T ˜ EC = n = 1 N W n TEC ( n ) ,
where T ~ EC is the reconstruction value, N is the total number of modeling stations, TEC(n) is the TEC observation value of the nth observation station, and Wn is the weight coefficients of the nth observation station. To ensure the unbiased reconstruction result, the weight coefficients should meet the following requirements:
n = 1 N W n = 1 ,
the weight coefficients are appropriately selected to make the reconstruction variance less than any linear combination of observation values. The minimum reconstruction variance is obtained by the appropriate variation function, whose equation is defined as:
γ ( d n j ) = 1 2 E [ ( TEC ( n ) TEC ( j ) ] 2 = 1 2 E [ ( TEC ( n ) TEC ( n + d n j ) ] 2 ,
where dnj is the ionospheric distance between the nth and jth station, TEC(n) is the observation at the nth station, and TEC(j) is the observation at the jth station whose distance is dnj from the nth station. According to (7), the variation function reveals the variation pattern in the entire spatial region, and the commonly used variation functions are Gaussian, exponential, sinusoidal, and linear models [67].
In conclusion, we can use the following equation to determine the optimal weight coefficients:
n = 1 N γ ( d n j ) × W n = γ ( d j 0 ) Φ   ,   j = 1 , 2 N ,
where dnj is the ionospheric distance between the nth and jth station, γ(dnj) is a variation function expressed as the ionospheric distance between the nth and jth station, dj0 is the ionospheric distance between the jth and reconstruction station, γ(dj0) is a variation function expressed as the ionospheric distance between the jth and reconstruction station, Wn is the weight coefficients of the nth station, N is the total number of modeling stations, and Φ is the Lagrange multiplier.
Combining (8) and the above, it can be seen that the variation function γ(dnj) of the TEC spatial reconstruction is the determination of the ionospheric distance between observation stations.
The spatial distance between two observed stations differs from the mathematically defined distance. The spatial resolution of the model can be accurate to 1 degree by multiplying 1 degree in geographic coordinates. Therefore, based on the linear variation model, this paper adopts the modified Euclidean distance with a scale factor (SF) as the ionospheric distance [68]; that is, the distance between any two observed stations can be expressed as:
d n j = ( φ n φ j ) 2 + [ S F ( λ n λ j ) ] 2 ,
where (φn,λn), and (φj,λj) represent the longitude and latitude coordinates of the nth and the jth ionospheric observed station, respectively; SF is the scale factor of correlation distance, and is related to the region located and reflects the correlation distance between ionospheric parameters in a region to some extent. Therefore, the goal of spatial reconstruction is to determine the optimal value of SF so that the reconstruction value is as close as possible to the observation value.
In the following subsection, we will discuss how to determine SF through cross-validation with RRMSE as the evaluation standard and then give the determined spatial reconstruction model.

2.4. Parameter Determination

2.4.1. Temporal Model Determination

Figure 8 shows the correlation coefficients between the observed and predicted TECs with F10.712, R12, and dual-parameter (F10.712+R12) regression for 24 h in the specific year of 2008, 2011, and 2014. It can be seen that the predicted TECs obtained by dual-parameter regression are highly correlated with the observed TECs compared with those obtained by single-parameter regression of F10.712 and R12. And the correlation coefficients of dual-parameter regression are closer to 1, which further confirms the correctness of introducing the dual-parameter into the model. In particular, the correlation coefficients are higher during periods of high solar activity than during other periods of solar activity, which has also been mentioned in [69,70].
Figure 9 shows the temporal regression results under different solar activity parameters, and the RMSEs and RRMSEs with F10.712, R12, and the dual-parameter. It is easy to see that the RRMSE of temporal regression with dual-parameter at the same time is less than that of single-parameter regression under the same conditions. Moreover, the regression results coincide with the TEC observations, in which some outliers were filtered out. Therefore, we simultaneously introduce F10.712 and R12 into the temporal regression model.
Figure 10a shows the RRMSE of TEC prediction when K and L take different values in the case of dual-parameter is introduced into the temporal map. When L is fixed, the value of K goes from 1 to 2 and then to 3, and the regression error decreases significantly, whereas K goes from 3 to 4, and the regression error does not change significantly. Therefore, we take K equal to 3 as the upper bound of the time cycle. When K is 3, the dual-parameter regression error corresponding to different L and the time required for regression are shown in Figure 10b. L increased from 3 to 5, the corresponding regression error decreased significantly, and regression time has not increased dramatically. In contrast, L rose from 5 to 7, the corresponding regression error decreased slightly, and the regression time increased sharply. Therefore, considering the regression error and regression time comprehensively, we take L equal to 5.
In conclusion, the temporal map (2) of TEC can be written as:
TEC = T ^ EC ( p , r , m , t )                   = k = 0 3 l = 0 5 [ c k , l p p l cos ( 2 π k m / 12 ) + s k , l p p l sin ( 2 π k m / 12 ) + ,                     c k , l r r l cos ( 2 π k m / 12 ) + s k , l r r l sin ( 2 π k m / 12 ) ] .
The undetermined coefficients in the temporal map can be determined by using the LS method with the RRMSE as the evaluation standard. Figure 11 shows the temporal regression coefficients of the Dourbes station as an example. It can be seen that the temporal regression coefficients increase significantly at noon when the solar radiation is the strongest compared with other times.

2.4.2. Spatial Model Determination

Considering the modeling data and spatial distribution of the modeling stations, we select the cross-validation of 8 modeling stations to determine the spatial reconstruction model. The specific process is shown in Figure 12. First, the TEC observations of the reconstruction station are removed from the modeling dataset. Then, the remaining observed TEC of the modeling stations (represented as the reconstruction dataset in Figure 12) were used to carry out spatial interpolation for the reconstruction stations, and the cycle was carried out successively until the TEC reconstruction of all stations was completed. And RRMSE was used as the evaluation standard to determine the optimal SF.
Figure 13 shows the RRMSE of the spatial reconstruction results against different SF. The minimum reconstruction error is 13.49%, and the corresponding SF is 1.15. Accordingly, the ionospheric distance between two ionospheric observation stations can be expressed as:
d n j = ( φ n φ j ) 2 + ( 1.15 ( λ n λ j ) 2 ) ,
where (φn,λn) and (φj,λj) represent the longitude and latitude coordinates of the nth and the jth ionospheric observation station, respectively.
After the definition of the ionospheric distance is determined, the Lagrange extreme algorithm can be applied to solve the variation function (8) and spatial interpolation (5), then the spatial reconstruction weight coefficients of each observation station can be determined. Figure 14 shows the spatial reconstruction weight coefficients corresponding to the 8 modeling stations used in this paper. It can be seen that different modeling stations have different weights on reconstructed TECs of the same reconstruction station. Specifically, the closer the spatial distance between the modeling station and the reconstruction station, the greater the corresponding weight coefficients of these stations.
To sum up, we determine the temporal and spatial reconstruction model of TEC, as well as the harmonic function composed by them. To validate the correctness and effectiveness of the proposed model, we will take RMSE and RRMSE as the evaluation standard to test the model by comparing the observed TECs of the validation stations and the predicted TECs calculated by using the empirical model determined by the above process. Furthermore, the proposed model is compared with the IRI model to prove its superiority.

3. Validation

As described in Section 2, the Nicosia, Chilton, and Sopron stations, which have less than 11 years of continuous data and are not involved in modeling, are used as validation stations. Meanwhile, to satisfy the comprehensiveness of validation, the Roquetes, San Vito, and Dourbes stations are selected from the modeling stations for validation. From the perspective of space, the geographical location distribution of validation stations is shown in Figure 15.
Considering the distribution of the modeling region, we took 10° east longitude as the dividing line, adhering to the principle of taking 3 stations for each side. From the perspective of time, the stations used for validation cover the four seasons of spring, summer, autumn, and winter in low and high solar activity years. The selection of validation stations fully ensures the non-overlapping of spatial and temporal distribution and the comprehensiveness of validation. The details are listed in Table 2.
For a given geographic latitude and longitude coordinates, year, month, universal time, F10.712, and R12, the TECs for any time and location can be predicted using the process shown in Figure 16. The specific process is as follows:
(1)
Using the temporal regression map (10) to calculate the time component T ^ EC .
(2)
Using the temporal regression value T ^ EC and the given geographical latitude and longitude coordinates, the final predicted T ~ EC value can be obtained by the modified Kriging interpolation method (5) for spatial reconstruction.
The predicted TECs of the 6 validation stations calculated by the proposed model (marked as Prop., black) are compared with the observed TECs (marked as Obs., blue) and the predicted TECs calculated by the IRI model with the CCIR coefficients (marked as CCIR, green) and the URSI coefficients (marked as URSI, red), respectively. Figure 17 shows the scatter plots of monthly mean values of TEC for a given year and month calculated using three models during three periods of solar activity.
To better evaluate the performance of the models, the statistical RMSEs and RRMSEs of predicted TECs for different models are listed in Table 3. At the end of Table 3, we give the average prediction error when using different models to predict any station; that is, the error of different models when predicting six validation stations is calculated again to obtain the average value.
Combined with Figure 17 and Table 3, we can draw the following four conclusions:
(1)
The TEC curve predicted by the IRI model and the model proposed in this paper has a highly fitting variation trend with the observed TEC curve. That is, there is an apparent variation trend of the day, month, season, and year as well as periodic variation with the solar activity. Furthermore, the maximum TEC value in high solar activity years is obviously higher than that in low solar activity years, which further confirms the close relationship between the variation of TEC and solar activity.
(2)
The prediction curve of the proposed model is closer to the TEC observation curve than that of the IRI model. Especially in Roquetes and Dourbes, the prediction effect of the proposed model is significantly better than that of the IRI model.
(3)
In Nicosia, the prediction error of both the IRI model and the proposed model is the maximum. The RRMSE of the IRI model using CCIR coefficients and URSI coefficients and the proposed model are 98.65%, 78.18%, and 21.71%, respectively, in predicting the monthly mean values of TEC of Nicosia station. The main reason is that the station is located at the junction of Asia and Europe, and the dynamic characteristics of the ionosphere are complex. In addition, the valid observation data provided by the observation station are insufficient, or the observation data provided by the limited ionospheric observation station are not accurate due to the limitation of observation technology, which significantly impacts the modeling accuracy.
(4)
When the IRI model was used to predict TEC of Sopron and Chilton stations, it could be found from the results that both CCIR coefficients and URSI coefficients showed double peaks that did not conform to the law of diurnal variation of TEC. This is one reason that the prediction error of the IRI model is higher than that of the proposed model.
To make the data more intuitive, Figure 18 shows the predicted RMSEs and RRMSEs of three models in six stations during three solar activity epochs, respectively. By comprehensively analyzing Figure 18 and Table 3, we can draw the following three conclusions:
(1)
Compared with the IRI model with CCIR and URSI coefficients, the predicted accuracy of the proposed model is improved by 38.63% and 35.79%, respectively, which is a significant improvement in parts of Europe. Similarly, other study [42] have also provided several sets of comparison graphs showing that the IRI model does have a significant deviation in TEC prediction, which further demonstrates the effectiveness of the PCA decomposition method in ionospheric parameters modeling and the necessity of the proposed model.
(2)
The prediction error of the proposed model is the minimum at the low solar activity epoch, which is 7.94%, and the maximum at the middle solar activity epoch, which is 17.75%, both of which are lower than the IRI models. On the other hand, the IRI model with CCIR and URSI coefficients has the best predictions during the high solar activity year, and the prediction errors are 26.34% and 31.48%, respectively.
(3)
The IRI model is an empirical model applicable to the global scope, whereas the proposed model is a regional empirical model focusing on the parts of Europe region. The prediction accuracy of the regional model is better than that of the global model in specific regions. Namely, the regional model has better spatial adaptability and pertinence in ionospheric parameter prediction.
The temporal resolution of the proposed model is 1h, and the spatial resolution can be accurate to 1° × 1° in geographic coordinates. Similarly, Zhang et al. used a 2.5°× 5° spatial grid and 1 h temporal resolution interpolated updated IG12 index to drive IRI-2016 to generate ionospheric TEC values consistent with IGS-TEC [17]. Figure 19 shows the spatial distribution of TECs predicted by the IRI model based on CCIR and URSI coefficients and the proposed model over parts of Europe at 06:00 UT in May 2009 (low solar activity year). It can be seen that TECs predicted by the three models show a similar spatial distribution. In particular, compared with the IRI model based on CCIR coefficients, the variation of the proposed model is closer to that of the IRI model based on URSI coefficients, which is consistent with the numerical validation results mentioned above. Furthermore, it can be seen that the distribution of TEC contour lines at high latitudes is relatively scattered; that is, the TECs in this region are relatively gentle. While the TEC contour lines at low latitudes are more concentrated, indicating that the TEC variations in this region are more intense. In addition, the numerical range of TECs predicted by the three models is basically the same, and the changes in details are slightly different. Compared with the sampling values of the modeling stations, the predicted value of the model proposed in this paper is closer to the sampling value at a given station, which proves the validity and correctness.
Figure 20 shows the spatial distribution of TECs predicted by the IRI model based on CCIR and URSI coefficients and the proposed model over parts of Europe at 06:00 UT in June 2014 (high solar activity year). In common with Figure 19, the TECs predicted by the three models show a similar spatial distribution; that is, the TECs located in the northwest region are smaller, while the TECs in the southeast region are larger. On the whole, the TECs predicted by the IRI model based on the URSI coefficients are the largest, followed by the IRI model based on the CCIR coefficients, and the predicted TECs of the model proposed in this paper are the smallest, which is consistent with the results mentioned in [42]; that is, when the IRI model predicts regional TEC, there will be an overall numerical deviation. Compared with the sampling values of the modeling stations, the predicted values of the proposed model are closer to the sampling values at a given station, and this model effectively improves the accuracy of the spatial TEC prediction model.

4. Conclusions

This paper proposes an ML-based modeling method and uses this method to establish an empirical prediction model of TEC for parts of Europe. It has been proved that the monthly mean values curve of TEC predicted by the model proposed in this paper is highly consistent with the observed curve of TEC. Furthermore, it is found that the prediction accuracy of the proposed model is significantly better than that of the IRI model with CCIR and URSI coefficients by comparing the observed TECs and the predicted TECs calculated by the IRI model and the proposed model, respectively. The improvement is 38.63% and 35.79%, respectively, which confirms the effectiveness of the proposed model.
This model is proposed for parts of Europe. Later we hope that it can have better effects and provide early warning for natural disasters and sudden magnetic storms by collecting more observed data, expanding the application scope, and improving the prediction accuracy. We will further explore the modeling methods for other regions to improve and enhance this model, which is compared with different models to validate the proposed model’s effectiveness and correctness based on more collected data such as GNSS TEC. In addition, we will finish the STFM of TEC based on this paper and make more efforts to improve the temporal and spatial resolution of the model and achieve a more accurate and detailed prediction of TEC. At the same time, we hope that this modeling idea can be applied to modeling other ionospheric parameters, such as the peak height of electron density in the F2 layer (hmF2), the maximum electron density in the F2 layer (NmF2), etc., and can provide a reference for modeling in other directions.

Author Contributions

Conceptualization, Y.L. and J.W.; methodology, Y.L. and J.W.; software, Y.L., C.Y. and J.W.; validation, Y.L., J.W. and H.F.; formal analysis, Y.L., C.Y. and J.W.; investigation, Y.L.; resources, J.W. and Y.Z; data curation, Y.L. and J.W.; writing—original draft preparation, Y.L. and J.W.; writing—review and editing, J.W., C.Y., Y.Z. and H.F; visualization, Y.L.; supervision, H.F.; project administration, J.W.; funding acquisition, C.Y. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 62031008) and the State Key Laboratory of Complex Electromagnetic Environment Effects on Electronics and Information System (No. CEMEE2022G0201, CEMEE-002-20220224).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yin, Y.; González-Casado, G.; Rovira-Garcia, A.; Juan, J.M.; Sanz, J.; Shao, Y. Summer Nighttime Anomalies of Ionospheric Electron Content at Midlatitudes: Comparing Years of Low and High Solar Activities Using Observations and Tidal/Planetary Wave Features. Remote Sens. 2022, 14, 1237. [Google Scholar] [CrossRef]
  2. Ren, X.; Liu, H.; Zhang, J.; Mei, D.; Zhang, X. An Improved Method for Ionospheric TEC Estimation Using the Spaceborne GNSS-R Observations. IEEE Trans. Geosci. Remote Sens. 2022, 60, 1–12. [Google Scholar] [CrossRef]
  3. Etemadfard, H.; Heinkelmann, R.; Hossainali, M.M.; Schuh, H. The Role of Spatial Gradient on Vertical Total Electron Content Extraction from Geodetic Very Long Baseline Interferometry Observation: Case Study CONT08 to CONT17-L1. Space Weather 2021, 19, e2020SW002633. [Google Scholar] [CrossRef]
  4. Wang, S.; Wang, D.; Sun, J. Artificial Neural Network-Based Ionospheric Delay Correction Method for Satellite-Based Augmentation Systems. Remote Sens. 2022, 14, 676. [Google Scholar] [CrossRef]
  5. Qiu, F.; Pan, X.; Luo, X.; Lai, Z.; Jiang, K.; Wang, G. Global ionospheric TEC prediction model integrated with semiparametric kernel estimation and autoregressive compensation. Chin. J. Geophys.-Chin. Ed. 2021, 64, 3021–3029. [Google Scholar] [CrossRef]
  6. Goss, A.; Hernández-Pajares, M.; Schmidt, M.; Roma-Dollase, D.; Erdogan, E.; Seitz, F. High-Resolution Ionosphere Corrections for Single-Frequency Positioning. Remote Sens. 2021, 13, 12. [Google Scholar] [CrossRef]
  7. Dong, Y.; Gao, C.; Long, F.; Yan, Y. Suspected Seismo-Ionospheric Anomalies before Three Major Earthquakes Detected by GIMs and GPS TEC of Permanent Stations. Remote Sens. 2022, 14, 20. [Google Scholar] [CrossRef]
  8. Zheng, Y.; Chernogor, L.F.; Garmash, K.P.; Guo, Q.; Rozumenko, V.T.; Luo, Y. Disturbances in the Ionosphere and Distortion of Radio Wave Characteristics That Accompanied the Super Typhoon Lekima Event of 4–12 August 2019. J. Geophys. Res. Space Phys. 2022, 127, 8. [Google Scholar] [CrossRef]
  9. Dao, T.; Harima, K.; Carter, B.; Currie, J.; McClusky, S.; Brown, R.; Rubinov, E.; Choy, S. Regional Ionospheric Corrections for High Accuracy GNSS Positioning. Remote Sens. 2022, 14, 2463. [Google Scholar] [CrossRef]
  10. He, J.; Yue, X.; Le, H.; Ren, Z.; Wan, W. Evaluation on the Quasi-Realistic Ionospheric Prediction Using an Ensemble Kalman Filter Data Assimilation Algorithm. Space Weather 2020, 18, e2019SW002410. [Google Scholar] [CrossRef]
  11. Tang, J.; Zhang, S.; Huo, X.; Wu, X. Ionospheric Assimilation of GNSS TEC into IRI Model Using a Local Ensemble Kalman Filter. Remote Sens. 2022, 14, 3267. [Google Scholar] [CrossRef]
  12. Farzaneh, S.; Forootan, E. Reconstructing Regional Ionospheric Electron Density: A Combined Spherical Slepian Function and Empirical Orthogonal Function Approach. Surv. Geophys. 2018, 39, 289–309. [Google Scholar] [CrossRef] [Green Version]
  13. Sharifi, M.A.; Farzaneh, S. The ionosphere electron density spatio-temporal modeling based on the Slepian basis functions. Acta Geod. Geophys. 2017, 52, 5–18. [Google Scholar] [CrossRef] [Green Version]
  14. Wang, L.; Wei, E.; Xiong, S.; Zhang, T.; Shen, Z. Evaluation of NeQuick2 Model over Mid-Latitudes of Northern Hemisphere. Remote Sens. 2022, 14, 4124. [Google Scholar] [CrossRef]
  15. Pietrella, M.; Pezzopane, M.; Zolesi, B.; Cander, L.R.; Pignalberi, A. The Simplified Ionospheric Regional Model (SIRM) for HF Prediction: Basic Theory, Its Evolution and Applications. Surv. Geophy. 2020, 41, 1143–1178. [Google Scholar] [CrossRef]
  16. Yao, Y.; Chen, X.; Kong, J.; Zhou, C.; Liu, L.; Shan, L.; Guo, Z. An Updated Experimental Model of IG12 Indices Over the Antarctic Region via the Assimilation of IRI2016 With GNSS TEC. IEEE Trans. Geosci. Remote Sens. 2021, 59, 1700–1717. [Google Scholar] [CrossRef]
  17. Zhang, W.; Huo, X.; Yuan, Y.; Li, Z.; Wang, N. Algorithm Research Using GNSS-TEC Data to Calibrate TEC Calculated by the IRI-2016 Model over China. Remote Sens. 2021, 13, 4002. [Google Scholar] [CrossRef]
  18. Bilitza, D. IRI the International Standard for the Ionosphere. Adv. Radio Sci. 2018, 16, 1–11. [Google Scholar] [CrossRef] [Green Version]
  19. Pignalberi, A.; Pezzopane, M.; Nava, B. Optimizing the NeQuick Topside Scale Height Parameters Through COSMIC/FORMOSAT-3 Radio Occultation Data. IEEE Geosci. Remote Sens. Lett. 2021, 19, 1–5. [Google Scholar] [CrossRef]
  20. Wang, J.; Liu, Y.; Xu, C. The Progress Review and Future Preview of Typical Ionospheric Models. In Proceedings of the International Symposium on Antennas, Propagation and EM Theory, Zhuhai, China, 30 November 2021. [Google Scholar]
  21. Kaselimi, M.; Voulodimos, A.; Doulamis, N.; Doulamis, A.; Delikaraoglou, D. Deep Recurrent Neural Networks for Ionospheric Variations Estimation Using GNSS Measurements. IEEE Trans. Geosci. Remote Sens. 2021, 60, 1–15. [Google Scholar] [CrossRef]
  22. Ren, X.; Yang, P.; Liu, H.; Chen, J.; Liu, W. Deep Learning for Global Ionospheric TEC Forecasting: Different Approaches and Validation. Space Weather 2022, 20, e2021SW003011. [Google Scholar] [CrossRef]
  23. Zhang, F.; Zhou, C.; Wang, C.; Zhao, J.; Liu, Y.; Xia, G.; Zhao, Z. Global ionospheric TEC prediction based on deep learning. Chin. J. Radio Sci. 2021, 36, 553–561. [Google Scholar]
  24. Sivakrishna, K.; Venkata Ratnam, D.; Sivavaraprasad, G. A Bidirectional Deep-Learning Algorithm to Forecast Regional Ionospheric TEC Maps. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2022, 15, 4531–4543. [Google Scholar] [CrossRef]
  25. Shi, S.; Zhang, K.; Wu, S.; Shi, J.; Hu, A.; Wu, H.; Li, Y. An Investigation of Ionospheric TEC Prediction Maps Over China Using Bidirectional Long Short-Term Memory Method. Space Weather 2022, 20, e2022SW003103. [Google Scholar] [CrossRef]
  26. Chen, Z.; Liao, W.; Li, H.; Wang, J.; Deng, X.; Hong, S. Prediction of Global Ionospheric TEC Based on Deep Learning. Space Weather 2022, 20, e2021SW002854. [Google Scholar] [CrossRef]
  27. Chen, J.; Fang, H.; Liu, Z. The Application of a Deep Convolutional Generative Adversarial Network on Completing Global TEC Maps. J. Geophys. Res. 2021, 126, e2020JA028418. [Google Scholar] [CrossRef]
  28. Pan, Y.; Jin, M.; Zhang, S.; Deng, Y. TEC Map Completion through a Deep Learning Model: SNP-GAN. Space Weather 2021, 11, e2021SW002810. [Google Scholar] [CrossRef]
  29. Lee, S.; Ji, E.; Moon, Y.; Park, E. One-Day Forecasting of Global TEC Using a Novel Deep Learning Model. Space Weather 2020, 19, 2020SW002600. [Google Scholar] [CrossRef]
  30. Ji, E.; Moon, Y.; Park, E. Improvement of IRI Global TEC Maps by Deep Learning Based on Conditional Generative Adversarial Networks. Space Weather 2020, 18, e2019SW002411. [Google Scholar] [CrossRef]
  31. Han, Y.; Wang, L.; Fu, W.; Zhou, H.; Li, T.; Chen, R. Machine Learning-Based Short-Term GPS TEC Forecasting during High Solar Activity and Magnetic Storm Periods. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2022, 15, 115–126. [Google Scholar] [CrossRef]
  32. Monte-Moreno, E.; Yang, H.; Hernández-Pajares, M. Forecast of the Global TEC by Nearest Neighbour Technique. Remote Sens. 2022, 14, 1361. [Google Scholar] [CrossRef]
  33. Tang, J.; Li, Y.; Yang, D.; Ding, M. An Approach for Predicting Global Ionospheric TEC Using Machine Learning. Remote Sens. 2022, 14, 1585. [Google Scholar] [CrossRef]
  34. Natras, R.; Soja, B.; Schmidt, M. Ensemble Machine Learning of Random Forest, AdaBoost and XGBoost for Vertical Total Electron Content Forecasting. Remote Sens. 2022, 14, 3547. [Google Scholar] [CrossRef]
  35. Mallika, I.L.; Ratnam, D.V.; Ostuka, Y.; Sivavaraprasad, G.; Raman, S. Implementation of Hybrid Ionospheric TEC Forecasting Algorithm Using PCA-NN Method. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2019, 12, 371–381. [Google Scholar] [CrossRef]
  36. Xiong, P.; Zhai, D.; Long, C.; Zhou, H.; Zhang, X.; Shen, X. Long Short-Term Memory Neural Network for Ionospheric Total Electron Content Forecasting Over China. Space Weather 2021, 19, e2020SW002706. [Google Scholar] [CrossRef]
  37. Jeong, S.; Lee, W.K.; Jang, S.; Kil, H.; Kim, J.; Kwak, Y.; Kim, Y.H.; Hong, J.; Choi, B. Reconstruction of the Regional Total Electron Content Maps over the Korean Peninsula Using Deep Convolutional Generative Adversarial Network and Poisson Blending. Space Weather 2022, 20, e2022SW003131. [Google Scholar] [CrossRef]
  38. Zewdie, G.K.; Valladares, C.; Cohen, M.B.; Lary, D.J.; Ramani, D.; Tsidu, G.M. Data-Driven Forecasting of Low-Latitude Ionospheric Total Electron Content Using the Random Forest and LSTM Machine Learning Methods. Space Weather 2021, 19, e2020SW002639. [Google Scholar] [CrossRef]
  39. Feng, J.; Wang, Z.; Jiang, W.; Zhao, Z.; Zhang, B. A single-station empirical model for TEC over the Antarctic Peninsula using GPS-TEC data. Radio Sci. 2017, 52, 196–214. [Google Scholar] [CrossRef]
  40. Krypiak-Gregorczyk, A.; Wielgosz, P.; Borkowski, A. Ionosphere Model for European Region Based on Multi-GNSS Data and TPS Interpolation. Remote Sens. 2017, 9, 2072–4292. [Google Scholar] [CrossRef] [Green Version]
  41. El-Diasty, M. Regional ionospheric modeling using wavelet network model. J. Glob. Position. Syst. 2017, 15, 1–10. [Google Scholar] [CrossRef] [Green Version]
  42. Basciftci, F.; Inal, C.; Yildirim, O.; Bülbül, S. Determining regional ionospheric model and comparing with global models. Geod. Vestn. 2017, 61, 427–440. [Google Scholar] [CrossRef]
  43. Nishioka, M.; Saito, S.; Tao, C.; Shiota, D.; Tsugawa, T.; Ishii, M. Statistical analysis of ionospheric total electron content (TEC): Long-term estimation of extreme TEC in Japan. Earth Planets Space 2021, 73, 52. [Google Scholar] [CrossRef]
  44. Fokoue, E.P. Model Selection for Optimal Prediction in Statistical Machine Learning. Not. Am. Math. Soc. 2020, 67, 155–168. [Google Scholar] [CrossRef]
  45. Shah, D.; Patel, K.; Shah, M. Prediction and estimation of solar radiation using artificial neural network (ANN) and fuzzy system: A comprehensive review. Int. J. Energ. Water Res. 2021, 5, 219–233. [Google Scholar] [CrossRef]
  46. Simon, H. Neural Networks and Learning Machines, 3rd ed.; Prentice Hall: Hoboken, NJ, USA, 2008. [Google Scholar]
  47. Xia, Z.; Chen, Y.; Xu, C. Multiview PCA: A Methodology of Feature Extraction and Dimension Reduction for High-Order Data. IEEE Trans. Cybern. 2021, 1–13. [Google Scholar] [CrossRef]
  48. Pearson, K. LIII. On lines and planes of closest fit to systems of points in space. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1901, 2, 559–572. [Google Scholar] [CrossRef] [Green Version]
  49. Dvinskikh, N.I. Expansion of ionospheric characteristics fields in empirical orthogonal functions. Adv. Space Res. 1988, 8, 179–187. [Google Scholar] [CrossRef]
  50. Wang, J.; Ma, J.G.; Huang, X.D.; Bai, H.M.; Chen, Q.; Cheng, H.Y. Modeling of the ionospheric critical frequency of the F2 layer over Asia based on modified temporal-spatial reconstruction. Radio Sci. 2019, 54, 680–691. [Google Scholar]
  51. Wang, J.; Yang, C.; An, W. Regional Refined Long-term Predictions Method of Usable Frequency for HF Communication Based on Machine Learning over Asia. IEEE Trans. Antennas Propag. 2022, 70, 4040–4055. [Google Scholar] [CrossRef]
  52. Ghaffari Razin, M.R.; Voosoghi, B. Ionosphere time series modeling using adaptive neuro-fuzzy inference system and principal component analysis. GPS Solut. 2020, 24, 1–13. [Google Scholar] [CrossRef]
  53. Wang, J.; Bai, H.M.; Huang, X.D.; Cao, Y.B.; Chen, Q.; Ma, J.G. Simplified regional prediction model of long-term trend for critical frequency of ionospheric F2 Region over east Asia. Appl. Sci. 2019, 9, 3219. [Google Scholar] [CrossRef] [Green Version]
  54. Galkin, I.; Froń, A.; Reinisch, B.; Hernández-Pajares, M.; Krankowski, A.; Nava, B.; Bilitza, D.; Kotulak, K.; Flisek, P.; Li, Z.; et al. Global Monitoring of Ionospheric Weather by GIRO and GNSS Data Fusion. Atmosphere 2022, 13, 371. [Google Scholar] [CrossRef]
  55. Space Weather Prediction Center. Available online: https://www.swpc.noaa.gov (accessed on 1 March 2021).
  56. Global Ionosphere Radio Observatory. Available online: https://giro.uml.edu/didbase/scaled.php (accessed on 30 November 2020).
  57. Wang, J.; Shi, Y.; Yang, C. Investigation of Two Prediction Models of Maximum Usable Frequency for HF Communication Based on Oblique- and Vertical-Incidence Sounding Data. Atmosphere 2022, 13, 1122. [Google Scholar] [CrossRef]
  58. Digisonde Portable Sounder 4 (DPS-4). Available online: http://digisonde.com/digisonde.html (accessed on 23 September 2022).
  59. Richards, P.G.; Fennelly, J.A.; Torr, D.G. EUVAC: A solar EUV Flux Model for aeronomic calculations. J. Geophys. Res. Space Phys. 1994, 99, 8981–8992. [Google Scholar] [CrossRef]
  60. De Haro Barbás, B.F.; Elias, A.G. Effect of the Inclusion of Solar Cycle 24 in the Calculation of foF2 Long-Term Trend for Two Japanese Ionospheric Stations. Pure Appl. Geophys. 2020, 177, 1071–1078. [Google Scholar] [CrossRef]
  61. Sivavaraprasad, G.; Deepika, V.S.; SreenivasaRao, D.; Ravi Kumar, M.; Sridhar, M. Performance evaluation of neural network TEC forecasting models over equatorial low-latitude Indian GNSS station. Geod. Geodyn. 2020, 11, 192–201. [Google Scholar] [CrossRef]
  62. Qian, L.; Burns, A.G.; Solomon, S.C.; Wang, W. Annual/semiannual variation of the ionosphere. Geophys. Res. Lett. 2013, 40, 1928–1933. [Google Scholar] [CrossRef]
  63. Bhawre, P.; Kishore, K.; Dogra, S.K.; Purohit, P.K.; Waheed, M.A.; Khatarkar, P.; Atulkar, R. Characteristic of Ionospheric foF2 and Solar Indices during the 23(rd) Solar Cycle over High Latitude Station, Syowa, Antarctica. Am. J. Clim. Chang. 2015, 4, 408–416. [Google Scholar] [CrossRef] [Green Version]
  64. Mustafa, U. Deep learning for ionospheric TEC forecasting at mid-latitude stations in Turkey. Acta Geophys. 2021, 60, 589–606. [Google Scholar]
  65. Zhukov, A.V.; Yasyukevich, Y.V.; Bykov, A.E. GIMLi: Global ionospheric total electron content model based on machine learning. GPS Solut. 2021, 25, 19. [Google Scholar] [CrossRef]
  66. Georges, M. Principles of geostatistics. Econ. Geol. 1963, 58, 1246–1266. [Google Scholar]
  67. Uddin, M.S.; Czajkowski, K.P. Performance Assessment of Spatial Interpolation Methods for the Estimation of Atmospheric Carbon Dioxide in the Wider Geographic Extent. J. Geovis. Spat. Anal. 2022, 6, 1–17. [Google Scholar] [CrossRef]
  68. Cheng, H.; Wang, J.; Huang, X. Reconstruction method for the ionospheric foF2 at high latitude region based on kriging method with geomagnetic coordinates. Chin. J. Space Sci. 2020, 40, 505–512. [Google Scholar]
  69. Ercha, A.; Zhang, D.-H.; Xiao, Z.; Hao, Y.-Q.; Ridley, A.J.; Moldwin, M. Modeling ionospheric foF2 by using empirical orthogonal function analysis. Ann. Geophys. 2011, 29, 1501–1515. [Google Scholar]
  70. Zhang, M.; Liu, C.; Wan, W.; Liu, L.; Ning, B. Evaluation of global modeling of M(3000)F2 and hmF2 based on alternative empirical orthogonal function expansions. Adv. Space Res. 2010, 46, 1024–1031. [Google Scholar] [CrossRef]
Figure 1. Four problems to be solved in ML and corresponding solutions.
Figure 1. Four problems to be solved in ML and corresponding solutions.
Remotesensing 14 05579 g001
Figure 2. Flowchart for TEC modeling and validation.
Figure 2. Flowchart for TEC modeling and validation.
Remotesensing 14 05579 g002
Figure 3. The geographical distribution of modeling and validation stations.
Figure 3. The geographical distribution of modeling and validation stations.
Remotesensing 14 05579 g003
Figure 4. Modeling and validation stations data and corresponding solar activity parameters: (a) the valid data period of the modeling and validation stations; (b) F10.712 and R12 variation curve.
Figure 4. Modeling and validation stations data and corresponding solar activity parameters: (a) the valid data period of the modeling and validation stations; (b) F10.712 and R12 variation curve.
Remotesensing 14 05579 g004
Figure 5. Comparison of monthly mean values of F10.712, R12, and TEC in the same period using the data of the Fairford station at UT 0:00 from 2004 to 2020: (a) F10.712; (b) R12; (c) TEC.
Figure 5. Comparison of monthly mean values of F10.712, R12, and TEC in the same period using the data of the Fairford station at UT 0:00 from 2004 to 2020: (a) F10.712; (b) R12; (c) TEC.
Remotesensing 14 05579 g005
Figure 6. The correlation coefficients between TEC and different solar activity parameters in 12 months and 24-h: (a) F10.712; (b) R12; (c) mean values of F10.712 and R12.
Figure 6. The correlation coefficients between TEC and different solar activity parameters in 12 months and 24-h: (a) F10.712; (b) R12; (c) mean values of F10.712 and R12.
Remotesensing 14 05579 g006
Figure 7. The flowchart of TEC spatial interpolation.
Figure 7. The flowchart of TEC spatial interpolation.
Remotesensing 14 05579 g007
Figure 8. The correlation coefficients between the observed and predicted TECs with F10.712, R12, and dual-parameter (F10.712+R12) regression for 24 h in the specific year of 2008, 2011, and 2014: (a) regression result using F10.712 in 2008; (b) regression result using R12 in 2008; (c) regression result using F10.712 and R12 in 2008; (d) regression result using F10.712 in 2011; (e) regression result using R12 in 2011; (f) regression result using F10.712 and R12 in 2011; (g) regression result using F10.712 in 2014; (h) regression result using R12 in 2014; (i) regression result using F10.712 and R12 in 2014.
Figure 8. The correlation coefficients between the observed and predicted TECs with F10.712, R12, and dual-parameter (F10.712+R12) regression for 24 h in the specific year of 2008, 2011, and 2014: (a) regression result using F10.712 in 2008; (b) regression result using R12 in 2008; (c) regression result using F10.712 and R12 in 2008; (d) regression result using F10.712 in 2011; (e) regression result using R12 in 2011; (f) regression result using F10.712 and R12 in 2011; (g) regression result using F10.712 in 2014; (h) regression result using R12 in 2014; (i) regression result using F10.712 and R12 in 2014.
Remotesensing 14 05579 g008
Figure 9. Regression results under different solar activity parameters: (a) comparison of TEC regression curves of the Roquetes station; (b) comparison of TEC regression curves of the Rome station; (c) TEC regression error of the Roquetes station; (d) TEC regression error of the Rome station.
Figure 9. Regression results under different solar activity parameters: (a) comparison of TEC regression curves of the Roquetes station; (b) comparison of TEC regression curves of the Rome station; (c) TEC regression error of the Roquetes station; (d) TEC regression error of the Rome station.
Remotesensing 14 05579 g009
Figure 10. Regression analysis results in the case of dual-parameter (F101.712+R12): (a) regression error against the upper bound of the time cycle; (b) regression error and modeling time corresponding to different L values when K is 3.
Figure 10. Regression analysis results in the case of dual-parameter (F101.712+R12): (a) regression error against the upper bound of the time cycle; (b) regression error and modeling time corresponding to different L values when K is 3.
Remotesensing 14 05579 g010
Figure 11. An example of temporal regression coefficients at the Dourbes station.
Figure 11. An example of temporal regression coefficients at the Dourbes station.
Remotesensing 14 05579 g011
Figure 12. The flowchart of spatial cross-validation.
Figure 12. The flowchart of spatial cross-validation.
Remotesensing 14 05579 g012
Figure 13. The RRMSE of spatial reconstruction against SF.
Figure 13. The RRMSE of spatial reconstruction against SF.
Remotesensing 14 05579 g013
Figure 14. The corresponding spatial reconstruction weight coefficients of the modeling stations.
Figure 14. The corresponding spatial reconstruction weight coefficients of the modeling stations.
Remotesensing 14 05579 g014
Figure 15. The geographical distribution of the validation stations.
Figure 15. The geographical distribution of the validation stations.
Remotesensing 14 05579 g015
Figure 16. Prediction flowchart.
Figure 16. Prediction flowchart.
Remotesensing 14 05579 g016
Figure 17. Day and night prediction results of TEC in different seasons: (a) San Vito station in May 2015; (b) Roquetes station in September 2014; (c) Dourbes station in October 2004; (d) Nicosia station in February 2016; (e) Sopron station in August 2020; (f) Chilton station in June 2009.
Figure 17. Day and night prediction results of TEC in different seasons: (a) San Vito station in May 2015; (b) Roquetes station in September 2014; (c) Dourbes station in October 2004; (d) Nicosia station in February 2016; (e) Sopron station in August 2020; (f) Chilton station in June 2009.
Remotesensing 14 05579 g017
Figure 18. The comparison of the TEC prediction results between the model proposed in this paper and the IRI model: (a) RMSE of different stations; (b) RRMSE of different stations; (c) RMSE of different solar activity; (d) RRMSE of different solar activity; (e) total RMSE; (f) total RRMSE.
Figure 18. The comparison of the TEC prediction results between the model proposed in this paper and the IRI model: (a) RMSE of different stations; (b) RRMSE of different stations; (c) RMSE of different solar activity; (d) RRMSE of different solar activity; (e) total RMSE; (f) total RRMSE.
Remotesensing 14 05579 g018
Figure 19. Spatial distribution of TECs predicted by CCIR, URSI and the proposed model in May 2009 (low solar activity year): (a) CCIR; (b) URSI; (c) Prop.
Figure 19. Spatial distribution of TECs predicted by CCIR, URSI and the proposed model in May 2009 (low solar activity year): (a) CCIR; (b) URSI; (c) Prop.
Remotesensing 14 05579 g019
Figure 20. Spatial distribution of TECs predicted by CCIR, URSI and the proposed model in June 2014 (high solar activity year): (a) CCIR; (b) URSI; (c) Prop.
Figure 20. Spatial distribution of TECs predicted by CCIR, URSI and the proposed model in June 2014 (high solar activity year): (a) CCIR; (b) URSI; (c) Prop.
Remotesensing 14 05579 g020
Table 1. Modeling and validation stations data information.
Table 1. Modeling and validation stations data information.
No.StationLatitude (°N)Longitude (°E)Duration of Valid DataData
Volume
Data SourcesReceiver
[57,58]
Including in Modeling
NOAA [55]GIRO [56]
1Athens38.0023.502004–20214948DPS-4Yes
2Chilton51.60−1.302005–20091417DPS-1No
3Dourbes50.104.602004–20215160DGS-256Yes
4El Arensillo37.10 −6.702004–20173834DGS-256Yes
5Fairford51.70−1.502004–20214726DGS-256Yes
6Nicosia35.1033.202013–20161546-DPS-4DNo
7Pruhonice50.0014.602004–20215136DPS-4Yes
8Rome41.8012.502006–20214584DPS-4Yes
9Roquetes41.000.002004–20215160DGS-256Yes
10San Vito40.6017.802004–20214779DGS-256Yes
11Sopron47.6316.722019–2021759-DPS-4DNo
Table 2. Validation stations data information.
Table 2. Validation stations data information.
No.StationLatitude (°N)Longitude (°E)Solar Activity EpochYearSeasonMonthIncluding in Modeling
1San Vito40.6017.80High2015SpringMayYes
2Roquetes41.000.00High2014SummerSeptemberYes
3Dourbes50.104.60Middle2004AutumnOctoberYes
4Nicosia35.1033.20Middle2016WinterFebruaryNo
5Sopron47.6316.72Low2020SummerAugustNo
6Chilton51.60−1.30Low2009SpringJuneNo
Table 3. RMSE and RRMSE of different models in TEC prediction.
Table 3. RMSE and RRMSE of different models in TEC prediction.
Statistical Analysis ItemRMSE (TECU)RRMSE (%)Difference between CCIR and Prop. (%)Difference between URSI and Prop. (%)
CCIRURSIProp.CCIRURSIProp.
StationSan Vito3.102.172.0018.5614.039.119.454.92
Roquetes3.804.632.1732.3042.2511.7420.5630.51
Dourbes3.683.541.0843.4544.2712.5830.8731.69
Nicosia4.923.862.4698.6578.1821.7176.9456.46
Sopron2.012.090.5142.3048.577.0435.2641.53
Chilton1.561.830.5332.3441.608.7423.6032.86
Solar
activity
High3.473.612.0826.3431.4810.5015.8420.98
Middle4.343.701.9076.2363.5317.7558.4845.78
Low1.801.970.5237.6545.227.9429.7137.28
Average3.373.201.6551.3948.5512.7638.6335.79
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Y.; Wang, J.; Yang, C.; Zheng, Y.; Fu, H. A Machine Learning-Based Method for Modeling TEC Regional Temporal-Spatial Map. Remote Sens. 2022, 14, 5579. https://doi.org/10.3390/rs14215579

AMA Style

Liu Y, Wang J, Yang C, Zheng Y, Fu H. A Machine Learning-Based Method for Modeling TEC Regional Temporal-Spatial Map. Remote Sensing. 2022; 14(21):5579. https://doi.org/10.3390/rs14215579

Chicago/Turabian Style

Liu, Yiran, Jian Wang, Cheng Yang, Yu Zheng, and Haipeng Fu. 2022. "A Machine Learning-Based Method for Modeling TEC Regional Temporal-Spatial Map" Remote Sensing 14, no. 21: 5579. https://doi.org/10.3390/rs14215579

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop