Home

Sources and Measurement of Traveling Ionospheric Disturbances

Introduction.

Traveling Ionospheric Disturbances (TIDs) are variations in the ionosphere that can impact medium frequency (MF) and high frequency (HF) radio communications through fading (QSB) and by causing variations in communications distance. HamSCI has two funded projects for studying TIDs and their connection to the atmosphere and space.

Science and Research Questions

  • Are TIDs with sources in the lower atmosphere are caused by secondary or tertiary AGWs that reach the thermosphere due to a multistep vertical coupling paradigm?
  • What  percentage of observed TIDs correlated with geomagnetic activity what percentage are not?
  • What is the TID longitudinal dependence on 2D stratospheric polar vortex configuration?
  • To what extent can data from amateur radio fill TID observational gaps and be scientifically useful?
  • To what extent will the use of advanced, physics-based whole atmospheric models in conjunction with observations improve our capacity to study the science of TIDs?

Publications

  • Frissell, N. A., Kaeppler, S. R., Sanchez, D. F., Perry, G. W., Engelke, W. D., Erickson, P. J., Coster, A. J., Ruohoniemi, J. M., Baker, J. B. H., West, M. lou. (2022). First Observations of Large Scale Traveling Ionospheric Disturbances Using Automated Amateur Radio Receiving Networks. Geophysical Research Letters (Under Review) . https://doi.org/https://doi.org/10.1002/essoar.10510169.1

Enabling Space Weather Research with Global Scale Amateur Radio Datasets

NASA SWO2R, 2 years (2021-2023), PI: Nathaniel Frissell W2NAF (University of Scranton), Co-Is: Travis Atkison (University of Alabama), William Engelke AB4EJ (University of Alabama), and Philip Erickson W1PJE (MIT Haystack Observatory)

  • Development of automated TID detection and parameter extraction algorithms.
  • Develop empirical TID models that use geophysical indices as independent variables and model the probability of TID occurrence signatures in terrestrial HF communications.
  • Validate models for the 7 and 14 MHz bands in the continental US and mainland Europe.
  • Deposit RBN/PSKReporter/WSPRNet data into public NASA data repositories.

CAREER: Amateur Radio as a Tool for Studying Traveling Ionospheric Disturbances and Atmosphere-Ionosphere Coupling

NSF CAREER, 5 years (2021-2026), PI: Nathaniel Frissell W2NAF (University of Scranton)

  • Identify the amount of TIDs observed by HF communications systems that are and are not associated with geomagnetic activity.
  • Determine the ability of data from amateur radio to fill TID observational gaps and be scientifically useful.
  • Establish TID longitudinal dependence on the 2D stratospheric polar vortex configuration.
  • Test the multistep vertical coupling paradigm of AGWs/TIDs theorized in the latest physics-based models.

travelling ionospheric disturbances

Article  

  • Volume 31, issue 2
  • ANGEO, 31, 163–172, 2013
  • Related articles

travelling ionospheric disturbances

GPS observations of medium-scale traveling ionospheric disturbances over Europe

S. nakagawa, m. nishioka, k. shiokawa.

Abstract. Two-dimensional structures of medium-scale traveling ionospheric disturbances (MSTIDs) over Europe have been revealed, for the first time, by using maps of the total electron content (TEC) obtained from more than 800 GPS receivers of the European GPS receiver networks. From statistical analysis of the TEC maps obtained 2008, we have found that the observed MSTIDs can be categorized into two groups: daytime MSTID and nighttime MSTID. The daytime MSTID frequently occurs in winter. Its maximum occurrence rate in monthly and hourly bin exceeds 70% at lower latitudes over Europe, whereas it is approximately 45% at higher latitudes. Since most of the daytime MSTIDs propagate southward, we speculate that they could be caused by atmospheric gravity waves in the thermosphere. The nighttime MSTIDs also frequently occur in winter but most of them propagate southwestward, in a direction consistent with the theory that polarization electric fields play an important role in generating the nighttime MSTIDs. The nighttime MSTID occurrence rate shows distinct latitudinal difference: The maximum of the occurrence rate in monthly and hourly bin is approximately 50% at lower latitudes in Europe, whereas the nighttime MSTID was rarely observed at higher latitudes. We have performed model calculations of the plasma density perturbations caused by a gravity wave and an oscillating electric field to reproduce the daytime and nighttime MSTIDs, respectively. We find that TEC perturbations caused by gravity waves do not show dip angle dependencies, while those caused by the oscillating electric field have a larger amplitude at lower latitudes. These dip angle dependencies of the TEC perturbation amplitude could contribute to the latitudinal variation of the MSTID occurrence rate. Comparing with previous studies, we discuss the longitudinal difference of the nighttime MSTID occurrence rate, along with the E- and F-region coupling processes. The seasonal variation, of the nighttime MSTID occurrence rate in Europe, is not consistent with the theory that the longitudinal and seasonal variations of the nighttime MSTID occurrence could be attributed to those of the Es layer occurrence.

  • Article (PDF, 2953 KB)
  • Supplement (18393 KB)
  • Article (2953 KB)
  • Metadata XML

Mendeley

  • Open access
  • Published: 20 January 2021

Solar activity dependence of medium-scale traveling ionospheric disturbances using GPS receivers in Japan

  • Yuichi Otsuka   ORCID: orcid.org/0000-0002-3098-3859 1 ,
  • Atsuki Shinbori 1 ,
  • Takuya Tsugawa 2 &
  • Michi Nishioka 2  

Earth, Planets and Space volume  73 , Article number:  22 ( 2021 ) Cite this article

3052 Accesses

22 Citations

2 Altmetric

Metrics details

In order to reveal solar activity dependence of the medium-scale traveling ionospheric disturbances (MSTIDs) at midlatitudes, total electron content (TEC) data obtained from a Global Positioning System (GPS) receiver network in Japan during 22 years from 1998 to 2019 were analyzed. We have calculated the detrended TEC by subtracting the 1-h running average from the original TEC data for each satellite and receiver pair, and made two-dimensional TEC maps of the detrended TEC with a spatial resolution of 0.15° × 0.15° in longitude and latitude. We have investigated MSTID activity, defined as \(\delta I/\overline{I}\) , where \(\delta I\) and \(\overline{I}\) are standard deviation of the detrended TEC and the average vertical TEC within the area of 133.0°–137.0° E and 33.0°–37.0° N for 1 h, respectively. From each 2-h time series of the detrended TEC data within the same area as the MSTID activity, auto-correlation functions (ACFs) of the detrended TEC were calculated to estimate the horizontal propagation velocity and direction of the MSTIDs. Statistical results of the MSTID activity and propagation direction of MSTIDs were consistent with previous studies and support the idea that daytime MSTIDs could be caused by atmospheric gravity waves, and that nighttime MSTIDs were caused by electro-dynamical forces, such as the Perkins instability. From the current long-term observations, we have found that the nighttime MSTID activity and occurrence rate increased with decreasing solar activity. For the daytime MSTID, the occurrence rate increased with decreasing solar activity, whereas the MSTID activity did not show distinct solar activity dependence. These results suggest that the secondary gravity waves generated by dissipation of the primary gravity waves propagating from below increase under low solar activity conditions. The mean horizontal phase velocity of the MSTIDs during nighttime did not show a distinct solar activity dependence, whereas that during daytime showed an anticorrelation with solar activity. The horizontal phase velocity of the daytime MSTIDs was widely distributed from 40 to 180 m/s under high solar activity conditions, whereas it ranged between 80 and 200 m/s, with a maximum occurrence at 130 m/s under low solar activity conditions, suggesting that gravity waves with low phase velocity could be dissipated by high viscosity in the thermosphere under low solar activity conditions.

travelling ionospheric disturbances

Introduction

In the Earth’s ionosphere, various types of disturbances of the plasma density exist. Traveling ionospheric disturbance (TID) is one of them, and is defined as wave-like phenomenon of the plasma density in the ionosphere. According to Hunsucker ( 1982 ), medium-scale TID (MSTID) is defined as a TID with horizontal wavelengths of approximately hundreds of kilometers, and periods ranging from 15 min to 1 h. Since Hines ( 1960 ), MSTIDs have been considered to be caused by atmospheric gravity waves (GWs) propagating in the thermosphere, and thus, a manifestation of the GWs. Although various techniques have been used to observe MSTIDs, most of the observations of the MSTIDs until the late 1990s were carried out at a single or a few points on the ground using ionosondes, high frequency (HF) radars, incoherent scatter radars, and beacon radio waves from satellites along the satellite paths (e.g., DuCastel and Faynot 1964 ; Morgan et al. 1978 ; Bristow and Greenwald 1995 ; Jacobson et al. 1995 ; Hocke et al. 1996 ; Oliver et al. 1997 ). Since the late 1990s, Global Positioning System (GPS) receiver networks and all-sky airglow imagers have made it possible to observe two-dimensional horizontal structures of the MSTIDs (e.g., Saito et al. 1998 ; Mendillo et al. 1997 ). As a result, a distinct difference in the MSTID propagation direction between daytime and nighttime has been revealed, suggesting that MSTIDs are generated by different mechanisms between daytime and nighttime (e.g., Miller et al. 1997 ; Kotake et al. 2007 ).

Since the nighttime MSTIDs propagated southwestward in the northern hemisphere and northwestward in the southern hemisphere (e.g., Shiokawa et al. 2003a ; Martinis et al. 2011 ), the Perkins instability (Perkins 1973 ) was considered to be a cause of the nighttime MSTIDs. Electric field perturbations associated with the nighttime MSTIDs have also been observed (Shiokawa et al. 2003b ). Based on these results, such this type of MSTID was named electrified MSTID (EMSTID) to differentiate from MSTIDs that could be caused by GWs (Narayanan et al. 2018 ).

Over Japan, Saito et al. ( 1998 ), for the first time, showed two-dimensional maps of TEC obtained from a dense GPS-receiver network in Japan for the first time. Otsuka et al. ( 2011 ) have analyzed the GPS-TEC data from one year in Japan, and showed seasonal and local time variations of MSTIDs. Using two-dimensional TEC map data obtained from Global Navigation Satellite System (GNSS) receiver networks, local time and seasonal variations of MSTIDs over South California (Kotake et al. 2007 ), Europe (Otsuka et al. 2013 ), Brazil (Figueiredo et al. 2018 ), and China (Ding et al. 2011 ) have been clarified statistically. However, the solar activity dependence of the MSTID characteristics has not been studied well. In this study, we have investigated MSTID characteristics statistically by analyzing 22 years of two-dimensional data of GPS-TEC over Japan, from 1998 to 2019. Based on the statistical results, we will discuss possible causes of the solar activity dependence of MSTIDs at midlatitudes.

Data and analysis method

Approximately 1300 dual-frequency (1.57542 GHz and 1.22760 GHz) GPS receivers have been installed in Japan by the Geospatial Information Authority of Japan as the GNSS Earth Observation Network System (GEONET) as of 2020. These GPS receivers provide carrier phase delays and group delays (C/A and/or P-code pseudoranges) of dual-frequency GPS signals every 30 s. The TEC along the ray path of the radio wave between the GPS satellite and receiver was accurately obtained from the carrier phase data. The ambiguity in the phase measurements due to the unknown initialization constant was corrected by the TEC obtained from the corresponding pseudorange data. The inter-frequency biases included in the TEC obtained from the above procedure were subtracted by using the method by Otsuka et al. ( 2002 ) to obtain absolute TEC. In this method, a weighted least-squares fitting was used.

TEC perturbations caused by MSTIDs were obtained by subtracting the 1-h running average (average over ± 30 min centered on the corresponding data) from the original TEC time series for each pair of satellites and receivers (Saito et al. 1998 ). TEC data with elevation angles larger than 35° were used in this study. The slant TEC was converted to vertical TEC by multiplying a slant factor defined as \(\tau_{0} /\tau_{1}\) , where \(\tau_{1}\) is the length of the ray path between 250 and 450 km altitudes, and \(\tau_{0}\) is the thickness of the ionosphere (200 km) for the zenith path. TEC was mapped on the ionospheric shell at an altitude of 300 km with a horizontal cell of 0.15° × 0.15° in latitude and longitude. The TEC data within each horizontal cell were then averaged. This high spatial resolution made it possible to detect the MSTID structure in the detrended TEC map. Figure  1 shows a two-dimensional map of detrended TEC over Japan at 15:00 UT (24:00 LT) on 8 June 2019. Wavefronts of the detrended TEC, elongating from northwest to southeast, can be seen. The method for deriving two-dimensional TEC maps was described in detail by Saito et al. ( 1998 ).

figure 1

Two-dimensional map of detrended TEC at 15:00 UT on 8 June (00:00 LT on 9 June) 2019. TEC data within an area enclosed by solid lines (133.0°–137.0° E and 33.0°–37.0° N) were used to estimate MSTID activity and horizontal propagation direction and velocity of MSTIDs

We have investigated the long-term variation of MSTID activity, defined as \(\delta I/\overline{I}\) , where \(\delta I\) and \(\overline{I}\) are the standard deviation of the detrended TEC and mean vertical TEC within an area of 133.0°–137.0° E and 33.0°–37.0° N during one hour, respectively, using the same method as Otsuka et al. ( 2008 ). The area used for the calculation of \(\delta I\) and \(\overline{I}\) is enclosed by the black lines in Fig.  1 .

In this study, we have newly developed an automated method to estimate the horizontal propagation direction and velocity of MSTIDs. First, we have calculated three-dimensional auto-correlation functions (ACFs) from 2-h time series of detrended TEC map data every 5 min within an area of 133.0°–137.0° E and 33.0°–37.0° N, which was the same as that for the estimation of the MSTID activity. Since MSTIDs observed with GPS-TEC in Japan have horizontal wavelengths ranging between 100 and 400 km (Otsuka et al. 2011 ), this area covers more than one horizontal wavelength of MSTIDs in Japan. Figure  2 shows two-dimensional ACFs as a function of lags in longitude and latitude at lag times of − 40, − 20, 0, and 20 min, respectively. In each panel, a high ACF region is found to be elongated along the northwest–southeast. This structure corresponds to the horizontal structure of TEC perturbations caused by MSTIDs with the wavefront elongating along the northwest–southeast. Maximum AFCs, which is shown by ‘ + ’ in each panel, appear at (lag number in longitude, lag number in latitude) = (5, 3), (2, 2), (0, 0), and (− 3, − 2) at the time lags of − 40, − 20, 0, and 20 min, respectively. This movement of the maximum ACF with time represents the propagation of the TEC perturbations. Latitudinal and longitudinal lag numbers of the maximum ACF at each time epoch were obtained for the lag time from − 50 to 50 min at an interval of 10 min. Figure  3 shows (a) latitudinal and (b) longitudinal lag numbers of maximum ACF at each time epoch as a function of lag time. The solid line in Fig.  3 represents a regression line of the latitudinal (longitudinal) lag number with respect to the lag time of the maximum ACF, assuming that the regression line passes the origin. To obtain the regression line, the inverse of the lag number in latitude (longitude) was used as a weighting function. In this process, the data with ACF exceeding 0.4 and MSTID activity exceeding 1% were used. The slope of the regression line represents the propagation velocity of the detrended TEC in the latitudinal and longitudinal directions, respectively. The negative slope in Fig.  3 a, b indicates southward and westward propagation, respectively. Since the cell size of the TEC map was 0.15° × 0.15° in latitude and longitude, a unit lag in latitude and longitude was approximately 16.6 km and 13.7 km, respectively. In this study, the regression line was determined from more than 4 lag data points. From these slopes, zonal and meridional propagation velocity, and direction of the TEC perturbations were obtained. For the case shown in Fig.  3 , the horizontal phase velocity and azimuth clockwise from due north were found to be 75.4 m/s and 239°, respectively. This method has been applied to the TEC data from 1998 to 2019. To exclude the effects of geomagnetic disturbances, the data with a Kp -index larger than 4 were excluded. We classified the obtained MSTID parameters into high and low solar activity conditions according to the solar activity index, which is the F 10.7 index. The high (low) solar activity conditions were defined as F 10.7 index larger than 120 (smaller than 110).

figure 2

Two-dimensional auto-correlation functions (ACFs) as a function of lag numbers in longitude and latitude at different lag time of a − 40, b − 20, c 0, and d 20 min, obtained from the detrended TEC within the area of 133.0°–137.0° E and 33.0°–37.0° N (shown in Fig.  1 ) and during a period of 2 h between 14:00 and 16:00 UT on 8 June 2019. A maximum ACF for each lag time is shown by a symbol ‘ + ’ in the figure

figure 3

a Latitudinal and b longitudinal lag numbers of the maximum ACF at each time epoch for the time lag from − 50 to 50 min at an interval of 10 min for the detrended TEC within the area of 133.0°–137.0° E and 33.0°–37.0° N between 14:00 and 16:00 UT on 8 June 2019. Solid line in each panel represents regression line of latitudinal (longitudinal) lag number with respect to the lag time of the maximum ACF, assuming that the regression line passes the origin. To obtain the regression line, inverse of lag number in latitude (longitude) was used as a weighting function

Figure  4 shows local time and seasonal variations of MSTID activity and MSTID occurrence rate under high and low solar activity conditions. The MSTID activity has been classified into monthly and hourly bin, and averaged in each bin. The MSTID occurrence rate is also classified into monthly and hourly bin. The MSTID occurrence was defined as the event that satisfies the following criteria: (1) MSTID activity was larger than 1%, and (2) horizontal propagation velocity of the TEC perturbations was determined. From the figures, both MSTID activity and MSTID occurrence rate were found to show similar local time and seasonal variations, and strongly dependent on the local time and season. Based on these statistical results, it was found that the daytime MSTID activity and occurrence rate were higher in winter than in the other seasons, and that the nighttime MSTID activity and occurrence rate displayed semiannual variation, with a primary peak in summer and a secondary peak in winter. Both MSTID activity and occurrence rate reached a peak simultaneously at 22–23 LT in nighttime summer. In nighttime winter, they were maximized at 23 LT under low solar activity conditions, and approximately 19 LT under high solar activity conditions. From the comparison between the results under high and low solar activity conditions, it was found that both MSTID activity and occurrence rate during daytime and nighttime were higher under low solar activity conditions than under high solar activity conditions. The occurrence rate of nighttime MSTID in summer was close to 100% under low solar activity condition. It was also found that the MSTID activity was higher during nighttime than during daytime.

figure 4

Local time and seasonal variations of a , b MSTID activity and c , d MSTID occurrence rate averaged within an area of 133.0°–137.0° E and 33.0°–37.0° N under high and low solar activity conditions

Figure  5 shows local time and seasonal variations of horizontal propagation direction and velocity of MSTID for high and low solar activity conditions. In the figures, the data are shown only when the MSTID occurrence rate in each cell exceeded 50%. It was found that propagation direction and velocity of MSTID depended on the local time. The MSTIDs propagated southward (approximately 180° in azimuth) during daytime and southwestward (approximately 240° in azimuth) during nighttime. The propagation velocity of MSTIDs was higher during daytime than during nighttime. The propagation direction of MSTIDs during both daytime and nighttime did not show a distinct solar activity dependence. The propagation velocity of MSTID during daytime tended to be larger under low solar activity conditions than under high solar activity conditions, whereas solar activity dependence of the propagation velocity during nighttime was not discernible. At dawn (03–06 LT), when MSTID activity and occurrence rate were high, westward propagation was seen in all seasons. Since the morning terminator moves westward, the enhancement of MSTID activity and occurrence rate with the westward propagation may be an effect of rapid increase in TEC at the sunrise due to the plasma production by solar radiation. In summer, after sunrise (06–09 LT), southeastward propagation at velocities of 70–130 m/s was observed. Otsuka et al. ( 2011 ), who have investigated GPS-TEC data by visual inspection, had shown that MSTID over Japan at dawn appeared frequently in summer and propagated mostly eastward. They had suggested that the dawn MSTID could be caused by atmospheric gravity waves and that the wind filtering mechanism could be operating effectively because the dawn MSTID propagated in the opposite direction to the background neutral winds. In this study, we focused on the daytime MSTIDs, which occur during 09–15 LT (00–06 UT) on January and December, and the nighttime MSTIDs, which occur during 19–01 LT (10–15 UT) between May and August.

figure 5

Local time and seasonal variations of a, b horizontal propagation direction and c , d velocity of MSTID in the area of 133.0°–137.0° E and 33.0°–37.0° N under high and low solar activity conditions. Black indicates missing data. The data are shown only when the MSTID occurrence rate in each cell exceeds 50%

Figure  6 shows the 22-year variation of (a) yearly mean F 10.7 index and (b and c) MSTID activity and occurrence rate in Japan during daytime and nighttime between 1998 and 2019. Distinct solar activity dependence of the MSTID activity and occurrence rate during nighttime can be seen. The nighttime MSTID activity increased with decreasing solar activity. On the other hand, the daytime MSTID activity did not show a distinct solar activity dependence, whereas the MSTID occurrence rate during daytime displayed anticorrelation with the solar activity. Figure  6 d, e shows the horizontal propagation direction and phase velocity of daytime and nighttime MSTIDs. The MSTIDs propagated southward during daytime and southwestward during nighttime. The propagation direction of the daytime MSTIDs did not depend on the solar activity, whereas that of the nighttime MSTIDs tended to be toward south of southwest during high solar activity period and west of southwest during low solar activity period. The horizontal phase velocity of daytime MSTIDs showed a distinct anticorrelation with solar activity, whereas that of nighttime MSTIDs did not. Mean phase velocity of daytime MSTIDs reached 140 m/s during low solar activity period, and was approximately 90 m/s during high solar activity period.

figure 6

Twenty-two year variations of yearly mean of a solar activity index, F 10.7 ; b MSTID activity; c MSTID occurrence rate; d and e horizontal propagation direction (azimuth clockwise from due north) and phase velocity of MSTID, respectively, between 1998 and 2019. Red and blue curves represent daytime (average over 09–15 LT between May and August) and nighttime (average over 19–01 LT on January and December) MSTIDs, respectively

The solar activity ( F 10.7 ) dependence of the daytime and nighttime MSTID activity was investigated quantitatively. In Fig.  7 , the daytime and nighttime MSTID activity for each year was an average of hourly MSTID activity during the daytime (09–15 LT) in winter (January and December) and an average over nighttime (19–01 LT) in summer (between May and August), respectively. Using the least-squares method, linear regression lines for the daytime and nighttime MSTID activity as a function of F 10.7 , respectively, were determined. It was found that the MSTID activity ( \(\delta I/\overline{I}\) ) can be expressed as a function of F 10.7 : \(\delta I/\overline{I} = 1.77 - 0.00238{ } \times F_{10.7}\) for daytime and \(\delta I/\overline{I} = 3.33 - { }0.00646{ } \times F_{10.7}\) for nighttime. Anticorrelation of the MSTID activity with the solar activity was clearly observed for both daytime and nighttime. The nighttime MSTID activity strongly depended on the solar activity, whereas the solar activity dependence of the daytime MSTID activity was weak.

figure 7

Solar activity ( F 10.7 ) dependence of the MSTID activity for daytime (cross) and nighttime (circle). The daytime and nighttime MSTID activity for each year is an average of hourly MSTID activity over daytime (09–15 LT) in winter (January and December) and average over nighttime (19–01 LT) in summer (between May and August, respectively. Solid lines represent linear regression of the daytime and nighttime MSTID activity as a function of F 10.7 , respectively

To investigate the solar activity dependence of the phase velocity for the daytime MSTIDs, the occurrence distribution of the phase velocity for high and low solar activity periods was investigated. Figure  8 shows the local time variation of the occurrence distribution in the horizontal phase velocity of the MSTIDs in winter (on January and December) under high solar activity conditions and under low solar activity conditions, respectively. For this study, the day on which the 81-day mean F 10.7 index was larger than 120 (smaller than 100) was classified into high (low) solar activity conditions. The horizontal phase velocity data are divided into 10-m/s bins. The occurrence rate shown in Fig.  8 was obtained in each hourly and 10-m/s bin. The horizontal phase velocity of the daytime MSTIDs was distributed mainly from 40 to 180 m/s under high solar activity conditions. On the other hand, under low solar activity conditions, the occurrence rate in the phase velocity was confined between 80 and 200 m/s, with a maximum occurrence at 130 m/s. It was found that the occurrence of the higher phase velocity (100–200 m/s) was larger under low solar activity conditions than under high solar activity conditions. The occurrence of a phase velocity of ~ 50 m/s after sunset (18–24 LT) under both high and low solar activity conditions represented the occurrence of nighttime MSTID in winter.

figure 8

Local time variation of horizontal phase velocity distribution of MSTIDs in winter (January and December) observed in Japan under a high and b low solar activity conditions. Occurrence rate of the phase velocity was obtained in each hourly and 10-m/s bin

From the statistical analysis of the MSTID activity and propagation direction for MSTIDs, we found that MSTIDs frequently propagated southward during daytime, and southwestward during nighttime. This feature is consistent with previous studies (e.g., Kotake et al. 2007 ; Otsuka et al. 2011 , 2013 ; Tsugawa et al. 2007 ). In this section, we focus on the solar activity dependnce of the MSTID activity and phase velocity.

MSTID activity

MSTIDs over Japan during the nighttime propagate mostly southwestward. The northwest–southeast alignment of the MSTID wavefront is a preferable condition for the Perkins instability (Perkins 1973 ). When the normal line to the wavefront of the plasma density perturbations exists between the east and the ionospheric current direction, the plasma density perturbations grow with time through the Perkins instability. Because of the southwestward preference of the nighttime MSTID propagation direction, the Perkins instability is considered to play an important role in generating the nighttime MSTIDs.

Both activity and occurrence rate of nighttime MSTID during 22 years showed clear anticorrelation with solar activity. They were high (low) for low (high) solar activity conditions. This feature was consistent with that reported by previous observations of the nighttime MSTIDs by GPS-TEC (e.g., Saito et al. 2002 ) and by 630.0-nm airglow images (e.g., Duly et al. 2013 ; Tsuchiya et al. 2018 ).

The anticorrelation of the MSTID activity and occurrence rate can be explained by the theory of the Perkins instability. The linear growth rate of the Perkins instability, \({\upgamma }\) , is inversely proportional to \(\left\langle {\nu_{in} } \right\rangle\) , where \(\left\langle {\nu_{in} } \right\rangle\) is the ion–neutral collision frequency ( \(\nu_{in}\) ) integrated along the geomagnetic field with a weighting function of the plasma density (Perkins 1973 ; Hamza 1999 ). Because \(\nu_{in}\) is proportional to the neutral density, \(\left\langle {\nu_{in} } \right\rangle\) increases with increasing solar activity, and thus \({\upgamma }\) shows anticorrelation with the solar activity.

At midlatitudes, the altitude of the F layer is supported by effective electric fields, \({\varvec{E}}^{*} = {\varvec{E}} + {\varvec{U}} \times {\varvec{B}}\) , against the downward gravity-induced diffusion velocity \({\raise0.7ex\hbox{$g$} \!\mathord{\left/ {\vphantom {g {\left\langle \nu \right\rangle }}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{${\left\langle \nu \right\rangle }$}}\) , where \({\varvec{E}}\) is the electric field, \({\varvec{U}}\) is the neutral wind, \({\varvec{B}}\) is the geomagnetic field, and \(g\) is the gravitational acceleration. Therefore, \({\upgamma }\) is also proportional to \({\varvec{E}}^{*}\) . On average, \({\varvec{U}}\) does not show distinct solar activity dependence (Emmert et al. 2006 ), but \({\varvec{E}}\) is smaller under low solar activity conditions than under high solar activity conditions (e.g., Fejer 1993 ; Takami et al. 1996 ). \({\varvec{E}}\) is a polarization electric field driven through the F -region dynamo (Rishbeth 1971 ) by \({\varvec{U}}\) . Through the F -region dynamo, \({\varvec{E}}\) is generated to satisfy \({\varvec{E}} + {\varvec{U}} \times {\varvec{B}} = 0\) , which satisfies the divergence-free conditions of the current, so that the direction of \({\varvec{E}}\) is opposite to \({\varvec{U}} \times {\varvec{B}}\) . Under high solar activity conditions, when the height-integrated Pedersen conductivity in the F region is significantly larger than that in the E region, \({\varvec{E}} + {\varvec{U}} \times {\varvec{B}}\) is nearly equal to zero. On the other hand, under low solar activity conditions, the height-integrated Pedersen conductivity in the F region is comparable to that in the E region so that \(\left| {\varvec{E}} \right|\) is smaller than \(\left| {{\varvec{U}} \times {\varvec{B}}} \right|\) . Consequently, \({\varvec{E}}^{*} = {\varvec{E}} + {\varvec{U}} \times {\varvec{B}}\) is larger under low solar activity conditions than under high solar activity conditions, so that the growth rate \({\upgamma }\) is larger under low solar activity conditions than under high solar activity conditions. Consequently, the solar activity dependence of the growth rate for the Perkins instability is consistent with that of the observed MSTID activity.

On the other hand, for the daytime MSTID, the occurrence rate showed anticorrelation of the solar activity, whereas the solar activity dependence of MSTID activity was weak. Since most of the daytime MSTIDs propagate equatorward, the daytime MSTIDs are thought to be caused by GWs propagating in the thermosphere. The plasma in the F region moves along the geomagnetic field lines through the ion–neutral collisions because the ion motion across the geomagnetic field line is restricted owing to the ion gyrofrequency being much higher than the ion–neutral collision frequency, and the current in the F region does not produce polarization electric fields due to the high conductivity in the daytime E region. GWs propagating equatorward and upward have larger amplitudes of neutral particle oscillation, parallel to the geomagnetic field lines, and thus induce larger plasma density perturbations in the ionosphere compared with GWs propagating in other directions (Hooke 1968 ). This directivity of the ion motion to the neutral particle oscillation caused by GWs could be responsible for the preference of the equatorward propagation for the daytime MSTIDs.

To maintain the kinetic energy of the upward propagating GWs, the decrease in the neutral density in the thermosphere causes an increase in the amplitude of the GWs. Bowman ( 1992 ) have suggested that the amplified GWs may be responsible for spread F, which is also a plasma density disturbance in the ionosphere. Due to this effect, the MSITD activity could increase with decreasing solar activity. However, the daytime MSTID activity obtained in this study did not show distinct solar activity dependence, suggesting that the daytime MSTID over Japan could not be caused by GWs propagating from the lower atmosphere directly. Vadas ( 2007 ) has shown that the effect of viscosity on the GWs increases with decreasing solar activity because the neutral density decreases with decreasing solar activity. Because the viscosity of the neutral atmosphere reduces amplitude of GWs, the MSTID activity could be suppressed under low solar activity conditions. The combination of these opposite effects may cause indistinct solar activity dependence of the daytime MSTID activity. On the other hand, owing to the high viscosity, GWs may dissipate and create horizontal body forces by deopsiting momentum into the background atmosphere, such that secondary GWs may be generated (Vadas and Liu 2009 ). The seocndary GWs could be generated frequently under low solar activity conditions because of the low neutral density. High occurrence of the secondary GWs under low solar activity conditoins may be responsible for the anticorrelation of the daytime MSTID occurrence rate with solar activity.

Phase velocity of MSTIDs

Daytime MSTID was found to occur most frequently in winter. This feature was consistent with MSTIDs in different regions (e.g., Kotake et al. 2006 ; Otsuka et al. 2011 ). Miyoshi et al. ( 2018 ) successfully reproduced MSTIDs generated by upward propagating GWs that are spontaneously generated in the global atmosphere–ionosphere coupled model, and showed that the daytime MSTID was more active in winter than in summer. According to the dispersion relation for GW, the vertical wave number ( \(m\) ) of GW is expressed as

where \(k\) and \(c\) are the horizontal wave number and horizontal phase speed of the gravity wave, respectively, \(N\) is the Brunt–Väisälä frequency, and \(H\) is the scale height of the neutral atmosphere. Background neutral wind ( \(U\) ) also affects the vertical wave number ( \(m\) ). From the dispersion relation, it was found that \(m\) was larger when \(\left( {c - U} \right)^{2}\) was smaller.

Neutral winds in the thermosphere are produced due to a balance between the pressure gradient force and ion drag. The pressure gradient forces are caused by the solar extreme ultraviolet (EUV) heating of the thermosphere. Therefore, the neutral winds in the thermosphere blow poleward during daytime and equatorward during nighttime. Regarding seasonal variation of the neutral winds in the thermosphere, the winds blow from the summer hemisphere to the winter hemisphere because the thermospheric temperature is higher in summer hemisphere than in winter hemisphere due to the difference in the solar zenith angle between summer and winter hemispheres. Consequently, the thermospheric winds during the daytime blow poleward, and are larger in winter than in summer. Such a feature can be seen in the seasonal variation of the thermospheric winds measured by the Middle and Upper atmosphere (MU) radar in Japan (Kawamura et al. 2000 ). For the GWs propagating equatorward, the background winds blow in the direction opposite to the GW propagation, and are larger in winter than in summer. Therefore, \(\left( {c - U} \right)^{2}\) is larger in winter than in summer. According to the dispersion relationship for GW, the vertical wave number is expected to be larger in summer than in winter, and hence, the vertical wavelength could be larger in winter than in summer. Due to the high viscosity and thermal conductivity in the thermosphere, GWs in the thermosphere are dissipated (Pitteway and Hines 1963 ). In particular, the GWs with shorter vertical wavelengths could likely be attenuated and dissipated, but the GWs with longer wavelengths could propagate for long distances while maintaining the amplitude of the oscillation. Considering the dissipation of GWs, it is expected that the GWs in the thermosphere could be attenuated in summer because the vertical wavelength is shorter, and could propagate for long distances in winter. This scenario can explain our statistical result, which shows that the daytime MSTIDs occur most frequently in winter.

Regarding solar activity dependence, our results showed that the average phase velocity of the daytime MSTIDs was larger under low solar activity conditions than under high solar activity conditions. This dependence of the phase velocity on solar activity is attributed to the dependence of distribution of the phase velocity occurrence rate on solar activity. Under high solar activity conditions, the phase velocities of daytime MSTID are distributed widely between 20 and 200 m/s, and their maximum occurrence is at 130 m/s. On the other hand, under low solar activity conditions, the phase velocities tend to be confined to a range between 100 and 200 m/s. The occurrence rate of the phase velocity between 20 and 100 m/s is lower under low solar activity conditions than under high solar activity conditions. The low occurrence rate of low phase velocity of daytime MSTIDs in the low solar activity could be explained in terms of the dissipation of GWs in the low solar activity. According to the dispersion relation for GW, the GWs with slow phase velocity (small c in Eq. ( 1 )) could have larger vertical wave number ( m ) or shorter vertical wavelength. The GWs with short vertical wavelengths are likely to dissipate, and only GWs with longer vertical wavelengths could survive in the thermosphere. Because viscosity is larger when the neutral density is small, the dissipation of GWs due to viscosity could occur under low solar activity conditions (Vadas 2007 ). Our results suggest that under low solar activity conditions, GWs with slow phase velocity tend to be attenuated in the thermosphere, and the daytime MSTIDs are mostly caused by GWs with higher phase velocity although the daytime MSTIDs under high solar activity are caused by GWs with various phase velocities up to approximately 200 m/s. From Fig.  8 , it can be seen that the occurrence rate of the higher phase velocity (100–200 m/s) increases with decreasing solar activity. This may be caused by an increase in the secondary GWs generated by dissipation of the primary GWs in the thermosphere, where the neutral density is low under low solar activity conditions, as discussed in the previous subsection.

Since GWs in the thermosphere are considered to propagate from the lower atmosphere, seasonal variation of GW activity in the thermosphere may be affected by that of GWs in the lower atmosphere. Murayama et al. ( 1994 ) reported annual variation of the GWs in the lower stratosphere, observed by the MU radar in Japan, and showed that GW kinetic energy is maximized in winter. They have suggested that the GW kinetic energy in the stratosphere over Japan is high when the subtropical jet stream at approximately 12-km altitude is strong. Satellite observations have revealed the global distribution of the GW activity in the stratosphere (Tsuda et al. 2000 ; Jiang et al. 2004 ), and have shown that the GW activity is highest in winter in both northern and southern hemispheres. Using the Atmospheric Infrared Sounder (AIRS) aboard NASA's Aqua satellite, Hoffmann et al. ( 2013 ), who have investigated the global distribution of hotspots of stratospheric GWs, have shown that most of the hotspots were related to the orographic GWs. High GW activity in the stratosphere in winter may be related to the high GW activity in the thermosphere, resulting in high MSTID activity during the daytime in winter. However, using the MU radar in Japan, Tsuda et al. ( 1990 ) reported that seasonal variation of GW kinetic energy in the mesosphere (65–85 km) showed maxima in both summer and winter, and minima during equinox. Further study of the vertical propagation of GWs from the lower atmosphere to the thermosphere may be needed to understand the underlying mechanisms for seasonal variation of the GWs in the thermosphere and daytime MSTIDs.

Conclusions

We have analyzed GPS-TEC data in Japan during 22 years from 1998 to 2019 to investigate the solar activity dependence of MSTID over Japan. Statistical results of the MSTID activity, MSTID occurrence rate, and propagation direction of the MSTIDs were consistent with the previous studies, suggesting that daytime MSTIDs could be caused by GWs, and that nighttime MSTIDs could be caused by the Perkins instability. The findings obtained in this study, based on the current long-term observations, are summarized as follows:

The MSTID activity defined as \(\delta I/\overline{I}\) , where \(\delta I\) and \(\overline{I}\) are the standard deviation of the detrended TEC and average vertical TEC during 1 h, respectively, and the occurrence rate of the nighttime MSTID increases with decreasing solar activity. This feature was consistent with that of the growth rate for the Perkins instability.

The occurrence rate of the daytime MSTID showed an anticorrelation with the solar activity, whereas the MSTID activity during daytime did not show distinct solar activity dependence. This result suggests that the daytime MSTID over Japan mostly could not be caused by GWs propagating from the lower atmosphere directly, but that the increase in the secondary GWs generated by dissipation of the primary GWs propagating from below, under low solar activity conditions, may contribute to the anticorrelation of daytime MSTID occurrence rate with solar activity.

The horizontal phase velocity of the nighttime MSITD did not distinctively depend on the solar activity. On the other hand, the mean horizontal phase velocity of the MSTIDs during daytime showed an anticorrelation with solar activity. The horizontal phase velocity of the daytime MSTIDs was widely distributed from 40 to 180 m/s widely, whereas it was confined to a range between 80 and 200 m/s with a maximum occurrence at 130 m/s under low solar activity conditions. This result suggests that GWs with low phase velocity could be dissipated by high viscosity in the thermosphere under low solar activity conditions.

Availability of data and materials

The GPS-TEC data are available at DRAWING-TEC ( https://aer-nc-web.nict.go.jp/GPS/DRAWING-TEC/ ).

Abbreviations

Auto-correlation function

Electrified medium-scale traveling ionospheric disturbance

Extreme ultraviolet

GNSS Earth Observation Network System

Global Navigation Satellite System

Global Positioning System

  • Gravity wave

High frequency

Medium-scale traveling ionospheric disturbance

  • Total electron content
  • Traveling ionospheric disturbance

Bowman GG (1992) Upper atmosphere neutral-particle density variations compared with spread-F occurrence rates at locations around the world. Ann Geophys 10:676–682

Google Scholar  

Bristow WA, Greenwald RA (1995) Estimating gravity wave parameters from oblique high-frequency backscatter: modeling and analysis. J Geophys Res 100(A3):3639–3648

Article   Google Scholar  

Ding F, Wan W, Xu G, Yu T, Yang G, Wang J (2011) Climatology of medium-scale traveling ionospheric disturbances observed by a GPS network in central China. J Geophys Res. https://doi.org/10.1029/2011JA016545

DuCastel F, Faynot JM (1964) Some irregularities observed simultaneously in the upper and lower ionosphere at middle latitudes. Nature 204:984

Duly TM, Chapagain NP, Makela JJ (2013) Climatology of nighttime medium-scale traveling ionospheric disturbances (MSTIDs) in the Central Pacific and South American sectors. Ann Geophys 31(12):2229–2237. https://doi.org/10.5194/angeo-31-2229-2013

Emmert JT, Faivre ML, Hernandez G, Jarvis MJ, Meriwether JW, Niciejewski RJ, Sipler DP, Tepley CA (2006) Climatologies of nighttime upper thermospheric winds measured by ground-based Fabry-Perot interferometers during geomagnetically quiet conditions: 1. Local time, latitudinal, seasonal, and solar cycle dependence. J Geophys Res. https://doi.org/10.1029/2006JA011948

Fejer BG (1993) F region plasma drifts over Arecibo: Solar cycle, seasonal, and magnetic activity effects. J Geophys Res 98(A8):13645–13652

Figueiredo CAOB, Takahashi H, Wrasse CM, Otsuka Y, Shiokawa K, Barros D (2018) Medium-scale traveling ionospheric disturbances observed by detrended total electron content maps over Brazil. J Geophys Res Space Phys 123:2215–2227. https://doi.org/10.1002/2017JA025021

Hamza AM (1999) Perkins instability revisited. J Geophys Res 104:22567–22575

Hines CO (1960) Internal atmospheric gravity waves at ionospheric heights. Can J Phys 38:1441–1481

Hocke K, Schlegel K, Kirchengast G (1996) Phases and amplitudes of TIDs in the high latitude F-region observed by EISCAT. J Atmos Terr Phys 58:245–255. https://doi.org/10.1016/0021-9169(95)00033-X

Hoffmann L, Xue X, Alexander MJ (2013) A global view of stratospheric gravity wave hotspots located with Atmospheric Infrared Sounder observations. J Geophys Res Atmos 118:416–434. https://doi.org/10.1029/2012JD018658

Hooke WH (1968) Ionospheric irregularities produced by internal atmospheric gravity waves. J Atmos Terr Phys 30:795–823

Hunsucker RD (1982) Atmospheric gravity waves generated in the high-latitude ionosphere: a review. Rev Geophys 20(2):293–315. https://doi.org/10.1029/RG020i002p00293

Jacobson AR, Carlos RC, Massey RS, Wu G (1995) Observation of traveling ionospheric disturbances with a satellite-beacon radio interferometer: seasonal and local time behavior. J Geophys Res 100(A2):1653–1665. https://doi.org/10.1029/94JA02663

Jiang JH, Wang B, Goya K, Hocke K, Eckermann SD, Ma J, Wu DL, Read WJ (2004) Geographical distribution and interseasonal variability of tropical deep convection: UARS MLS observations and analyses. J Geophys Res 109:D03111. https://doi.org/10.1029/2003JD003756

Kawamura S, Otsuka Y, Zhang SR, Fukao S, Oliver WL (2000) A climatology of middle and upper atmosphere radar observations of thermospheric winds. J Geophys Res 105(A6):12777–12788. https://doi.org/10.1029/2000JA900013

Kotake N, Otsuka Y, Tsugawa T, Ogawa T, Saito A (2006) Climatological study of GPS total electron content variations caused by medium-scale traveling ionospheric disturbances. J Geophys Res. https://doi.org/10.1029/2005JA011418

Kotake N, Otsuka Y, Tsugawa T, Ogawa T, Saito A (2007) Statistical study of medium-scale traveling ionospheric disturbances observed with the GPS networks in Southern California. Earth Planets Space 59:95–102. https://doi.org/10.1186/BF03352681

Martinis C, Baumgardner J, Wroten J, Mendillo M (2011) All-sky imaging observations of conjugate medium-scale traveling ionospheric disturbances in the American sector. J Geophys Res. https://doi.org/10.1029/2010JA016264

Mendillo M, Baumgardner J, Nottingham D, Aarons J, Reinisch B, Scali J, Kelley M (1997) Investigations of thermospheric-ionospheric dynamics with 6300 a images from the Arecibo Observatory. J Geophys Res 102:7331–7343

Miller CA, Swartz WE, Kelley MC, Mendillo M, Nottingham D, Scali J, Reinisch B (1997) Electrodynamics of midlatitude spread F, 1. Observations of unstable, gravity wave-induced ionospheric electric fields at tropical latitudes. J Geophys Res 102:11521–11532

Miyoshi Y, Jin H, Fujiwara H, Shinagawa H (2018) Numerical study of traveling ionospheric disturbances generated by an upward propagating gravity wave. J Geophys Res Space Phys 123:2141–2155. https://doi.org/10.1002/2017JA025110

Morgan MG, Calderón CHJ, Ballard KA (1978) Techniques for the study of TIDs with multi-station rapid-run ionosondes. Radio Sci 13:729–741. https://doi.org/10.1029/RS013i004p00729

Murayama Y, Tsuda T, Kato S, Fukao S (1994) Seasonal variation of gravity wave activity in the lower stratosphere observed with the MU radar. J Geophys Res 99:23057–23069

Narayanan VL, Shiokawa K, Otsuka Y, Neudegg D (2018) On the role of thermospheric winds and sporadic E layers in the formation and evolution of electrified MSTIDs in geomagnetic conjugate regions. J Geophys Res Space Phys. https://doi.org/10.1029/2018JA025261

Oliver WL, Otsuka Y, Sato M, Takami T, Fukao S (1997) A climatology of F region gravity wave propagation over the middle and upper atmosphere radar. J Geophys Res 102(A7):14499–14512. https://doi.org/10.1029/97JA00491

Otsuka Y, Ogawa T, Saito A, Tsugawa T, Fukao S, Miyazaki S (2002) A new technique for mapping of total electron content using GPS network in Japan. Earth Planets Space 54:63–70. https://doi.org/10.1186/BF03352422

Otsuka Y, Tani T, Tsugawa T, Ogawa T (2008) Statistical study of relationship between medium-scale traveling ionospheric disturbance and sporadic E layer activities in summer night over Japan. J Atmos Solar-Terr Phys 70:2196–2202

Otsuka Y, Kotake N, Shiokawa K, Ogawa T, Tsugawa T, Saito A (2011) Statistical study of medium-scale traveling ionospheric disturbances observed with a GPS receiver network in Japan. In: Abdu M, Pancheva D, Bhattacharyya A (eds) Aeronomy of the Earth’s atmosphere and ionosphere. Springer, Netherlands

Otsuka Y, Suzuki K, Nakagawa S, Nishioka M, Shiokawa K, Tsugawa T (2013) GPS observations of medium-scale traveling ionospheric disturbances over Europe. Ann Geophys 31(2):163–172. https://doi.org/10.5194/angeo-31-163-2013

Perkins F (1973) Spread F and ionospheric currents. J Geophys Res 78:218–226

Pitteway MLV, Hines CO (1963) The viscous damping of atmospheric gravity waves. Can J Phys 41:1935–1948

Rishbeth H (1971) The F-layer dynamo. Planet Space Sci 19(2):263–267. https://doi.org/10.1016/0032-0633(71)90205-4

Saito A, Miyazaki S, Fukao S (1998) High resolution mapping of TEC perturbations with the GSI GPS network over Japan. Geophys Res Lett 25:3079–3082

Saito A, Nishimura M, Yamamoto M, Fukao S, Tsugawa T, Otsuka Y, Miyazaki S, Kelley MC (2002) Observations of traveling ionospheric disturbances and 3-m scale irregularities in the nighttime F-region ionosphere with the MU radar and a GPS network. Earth Planets Space 54:31–44. https://doi.org/10.1186/BF03352419

Shiokawa K, Ihara C, Otsuka Y, Ogawa T (2003a) Statistical study of nighttime medium-scale traveling ionospheric disturbances using midlatitude airglow imagers. J Geophys Res 108(A1):1052. https://doi.org/10.1029/2002JA009491

Shiokawa K, Otsuka Y, Ihara C, Ogawa T, Rich FJ (2003b) Ground and satellite observations of nighttime medium-scale traveling ionospheric disturbance at midlatitude. J Geophys Res 108(A4):1145. https://doi.org/10.1029/2002JA009639

Takami T, Oliver WL, Richmond AD, Fukao S (1996) Ionospheric drift similarities at magnetic conjugate and nonconjugate locations. J Geophys Res 101(A7):15773–15782. https://doi.org/10.1029/96JA01316

Tsuchiya S, Shiokawa K, Fujinami H, Otsuka Y, Nakamura T, Yamamoto M (2018) Statistical analysis of the phase velocity distribution of mesospheric and ionospheric waves observed in airglow images over a 16-year period: Comparison between Rikubetsu and Shigaraki. Japan J Geophys Res Space Phys. https://doi.org/10.1029/2018JA025585

Tsuda T, Murayama Y, Yamamoto M, Kato S, Fukao S (1990) Seasonal variation of momentum flux in the mesosphere observed with the MU radar. Geophys Res Lett 17:725–728. https://doi.org/10.1029/GL017i006p00725

Tsuda T, Nishida M, Rocken C, Ware RH (2000) A global morphology of gravity wave activity in the stratosphere revealed by the GPS occultation data (GPS/MET). J Geophys Res 105:7257–7273. https://doi.org/10.1029/1999JD901005

Tsugawa T, Kotake N, Otsuka Y, Saito A (2007) Medium-scale traveling ionospheric disturbances observed by GPS receiver network in Japan: a short review. GPS Solut 11:139–144. https://doi.org/10.1007/s10291-006-0045-5

Vadas SL (2007) Horizontal and vertical propagation and dissipation of gravity waves in the thermosphere from lower atmospheric and thermospheric sources. J Geophys Res. https://doi.org/10.1029/2006JA011845

Vadas SL, Liu H (2009) Generation of large-scale gravity waves and neutral winds in the thermosphere from the dissipation of convectively generated gravity waves. J Geophys Res. https://doi.org/10.1029/2009JA014108

Download references

Acknowledgements

The GPS data in Japan were provided by The Geospatial Information Authority of Japan. The GPS-TEC data are archived by the National Institute of Information and Communications Technology, Japan. Geomagnetic and solar activity indices were obtained from the GSFC/SPDF OMNIWeb ( https://omniweb.gsfc.nasa.gov ).

The present study was supported by the Japan Society for the Promotion of Science, KAKENHI (15H05815, 16H06286), Project for Solar-Terrestrial Environment Prediction (PSTEP) and Study of dynamical variation of particles and waves in the inner magnetosphere using ground-based network observations (PWING).

Author information

Authors and affiliations.

Institute for Space-Earth Environmental Research, Nagoya University, Nagoya, Japan

Yuichi Otsuka & Atsuki Shinbori

National Institute of Information and Communications Technology, Tokyo, Japan

Takuya Tsugawa & Michi Nishioka

You can also search for this author in PubMed   Google Scholar

Contributions

YO designed the study, performed the data analysis, and prepared the manuscript. AS developed the GPS-TEC database. TT and MN processed the GPS data, and provided GPS-TEC data. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Yuichi Otsuka .

Ethics declarations

Ethics approval and consent to participate.

Not applicable.

Consent for publication

Competing interests.

The authors declare that they have no competing interests.

Additional information

Publisher's note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Otsuka, Y., Shinbori, A., Tsugawa, T. et al. Solar activity dependence of medium-scale traveling ionospheric disturbances using GPS receivers in Japan. Earth Planets Space 73 , 22 (2021). https://doi.org/10.1186/s40623-020-01353-5

Download citation

Received : 09 September 2020

Accepted : 31 December 2020

Published : 20 January 2021

DOI : https://doi.org/10.1186/s40623-020-01353-5

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Perkins instability

travelling ionospheric disturbances

A general climatology of categorized Martian ionospheric irregularities

  • Published: 04 September 2024

Cite this article

travelling ionospheric disturbances

  • Xin Wan 1 , 2 ,
  • Jiahao Zhong 1 , 2 ,
  • Yongqiang Hao 1 ,
  • Chao Xiong 3 ,
  • Hui Wang 3 ,
  • Yutian Cao 1 ,
  • Jun Cui 1 ,
  • Yiwen Liu 4 ,
  • Qiaoling Li 5 &
  • Jiawei Kuai 6  

A climatological survey of Martian ionospheric plasma density irregularities was conducted by exploring the in-situ measurements of the Mars Atmosphere and Volatile EvolutioN (MAVEN) spacecraft. The irregularities were first classified as enhancement, depletion, and oscillation. By checking the simultaneous magnetic field fluctuation, the irregularities have been classified into two types: with or without magnetic signatures. The classified irregularities exhibit diverse global occurrence patterns, as those with magnetic signatures tend to appear near the periphery of the crustal magnetic anomaly (MA), and those without magnetic signatures prefer to appear either inside of the MA or outside of the MA, depending on the type and solar zenith angle. Under most circumstances, the irregularities have a considerable occurrence rate at altitudes above the ionospheric dynamo height (above 200 km), and the magnetization state of the ions seems irrelevant to their occurrence. In addition, the irregularities do not show dependence on magnetic field geometry, except that the enhancement without magnetic signatures favors the vertical field line, implying its equivalence to the localized bulge. Other similarities and discrepancies exist in reference to previous studies. We believe this global survey complements previous research and provides crucial research clues for future efforts to clarify the nature of the Martian ionospheric irregularities.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

Abdu M, Kherani E A. 2011. Coupling processes in the equatorial spread F/plasma bubble irregularity development. In: Abdu M A, Pancheva D, Bhattacharyya A, eds. Aeronomy of the Earth’s Atmosphere and Ionosphere, IAGA Special Sopron Book Series 2. Springer Science +Business Media B. V. Chap. 16. 219–238

Chapter   Google Scholar  

Acuña M H, Connerney J E, Ness N F, Lin R P, Mitchell D, Carlson C W, McFadden J, Anderson K, Reme H, Mazelle C. 1999. Global distribution of crustal magnetization discovered by the mars global surveyor mag/er experiment. Science, 284: 790–793

Article   Google Scholar  

Alken P, Maute A, Richmond A D. 2017. The F-region gravity and pressure gradient current systems: A review. Space Sci Rev, 206: 451–469

Andersson L, Ergun R E, Delory G T, Eriksson A, Westfall J, Reed H, McCauly J, Summers D, Meyers D. 2015. The Langmuir probe and waves (LPW) instrument for MAVEN. Space Sci Rev, 195: 173–198

Andrews D J, André M, Opgenoorth H J, Edberg N J, Diéval C, Duru F, Gurnett D A, Morgan D, Witasse O. 2014. Oblique reflections in the Mars Express MARSIS data set: Stable density structures in the Martian ionosphere. J Geophys Res-Space Phys, 119: 3944–3960

Basuvaraj P, Němec F, Němeček Z, Šafránková J. 2022. Ionospheric plasma depletions at Mars observed by the MAVEN spacecraft. J Geophys Res-Planets, 127: e2022JE007302

Booker H G, Wells H W. 1938. Scattering of radio waves by the F-region of the ionosphere. J Geophys Res, 43: 249–256

Google Scholar  

Cao Y, Niu D, Liang W, Cui J, Wu X, Wan X, Zhong J, Li L, Rong Z, Wei Y. 2023. Total electron content depression in the nightside Martian ionosphere: Statistical results from Mars Express MARSIS measurements and implications. Mon Not R Astron Soc, 519: 2262–2267

Chicarro A, P Martin, Trautner R. 2004. The Mars express mission: An overview. In: Wilson A, ed. Mars Express: A European Mission to the Red Planet. Noordwijk, Netherlands: ESA Publ. Div. 3–16

Collinson G A, McFadden J, Grebowsky J, Mitchell D, Lillis R, Withers P, Vogt M F, Benna M, Espley J, Jakosky B. 2020. Constantly forming sporadic E-like layers and rifts in the Martian ionosphere and their implications for Earth. Nat Astron, 4: 486–491

Collinson G, McFadden J, Mitchell D, Grebowsky J, Benna M, Espley J, Jakosky B. 2019. Traveling ionospheric disturbances at Mars. Geophys Res Lett, 46: 4554–4563

Connerney J E P, Espley J, Lawton P, Murphy S, Odom J, Oliversen R, Sheppard D. 2015. The MAVEN magnetic field investigation. Space Sci Rev, 195: 257–291

Cravens T E, Fowler C M, Brain D, Rahmati A, Xu S, Ledvina S A, Andersson L, Renzaglia A R. 2020. Magnetic reconnection in the ionosphere of Mars: The role of collisions. J Geophys Res-Space Phys, 125: e2020JA028036

David M, Sojka J J, Schunk R W, Coster A J. 2016. Polar cap patches and the tongue of ionization: A survey of GPS TEC maps from 2009 to 2015. Geophys Res Lett, 43: 2422–2428

Diéval C, Morgan D D, Němec F, Gurnett D A. 2014. MARSIS observations of the Martian nightside ionosphere dependence on solar wind conditions. J Geophys Res-Space Phys, 119: 4077–4093

Diéval C, Andrews D J, Morgan D D, Brain D A, Gurnett D A. 2015. MARSIS remote sounding of localized density structures in the dayside Martian ionosphere: A study of controlling parameters. J Geophys Res-Space Phys, 120: 8125–8145

Diéval C, Kopf A J, Wild J A. 2018. Shapes of magnetically controlled electron density structures in the dayside Martian ionosphere. J Geophys Res-Space Phys, 123: 3919–3942

Duru F, Gurnett D A, Averkamp T F, Kirchner D L, Huff R L, Persoon A M, Plaut J J, Picardi G. 2006. Magnetically controlled structures in the ionosphere of Mars. J Geophys Res, 111: A12204

Foster J C, Zou S, Heelis R A, Erickson P J. 2021. Ionospheric storm-enhanced density plumes. In: Huang C, Lu G, Zhang Y, Paxton L J, eds. Ionosphere Dynamics and Applications. New York: John Wiley & Sons, Inc. 115–126

Fowler C M, Andersson L, Shaver S R, Thayer J P, Huba J D, Lillis R, Usanova M E, Espley J, Ergun R E, Mcfadden J, Mahaffy P R, Connerney J E P, Benna M, Elrod M, Mitchell D L, Mazelle C, Jakosky B M. 2017. MAVEN observations of ionospheric irregularities at Mars. Geophys Res Lett, 44: 10845–10854

Fowler C M, Bonnell J W, Huba J D, Andersson L, Espley J, Benna M, Ergun R E. 2019. The statistical characteristics of small-scale ionospheric irregularities observed in the Martian ionosphere. J Geophys Res-Space Phys, 124: 5874–5893

Fowler C M, Bonnell J W, Xu S, Benna M, Elrod M, McFadden J, Mitchell D L, Espley J, Andersson L, Ergun R E, Lillis R, Jakosky B. 2020. First detection of kilometer-scale density irregularities in the Martian ionosphere. Geophys Res Lett, 47: e2020GL090906

Fritts D C, Wang L, Tolson R H. 2006. Mean and gravity wave structures and variability in the Mars upper atmosphere inferred from Mars Global Surveyor and Mars Odyssey aerobraking densities. J Geophys Res, 111: A12304

Gurnett D A, Kirchner D L, Huff R L, Morgan D D, Persoon A M, Averkamp T F, Duru F, Nielsen E, Safaeinili A, Plaut J J, Picardi G. 2005. Radar soundings of the ionosphere of Mars. Science, 310: 1929–1933

Article   CAS   Google Scholar  

Gurnett D A, Morgan D D, Duru F, Akalin F, Winningham J D, Frahm R A, Dubinin E, Barabash S. 2010. Large density fluctuations in the Martian ionosphere as observed by the Mars Express radar sounder. Icarus, 206: 83–94

Haldoupis C. 2012. Midlatitude sporadic E: A typical paradigm of atmosphere-ionosphere coupling. Space Sci Rev, 168: 441–461

Harada Y, Gurnett D A, Kopf A J, Halekas J S, Ruhunusiri S. 2018. Ionospheric irregularities at Mars probed by MARSIS topside sounding. J Geophys Res-Space Phys, 123: 1018–1030

Hines C O. 1964. The formation of midlatitude sporadic E layers. J Geophys Res, 69: 1018–1019

Hinson D P, Tyler G L, Hollingsworth J L, Wilson R J. 2001. Radio occultation measurements of forced atmospheric waves on Mars. J Geophys Res, 106: 1463–1480

Huang J, Wu Z, Cui J, Hao Y, Cao W. 2022. Topographic gravity waves observed in the Martian thermosphere: A statistical perspective from a 1-D full-wave model. J Geophys Res-Space Phys, 127: e2022JA030846

Jakosky B M, Lin R P, Grebowsky J M, Luhmann J G, Mitchell D F, Beutelschies G, Priser T, Acuna M, Andersson L, Baird D, Baker D, Bartlett R, Benna M, Bougher S, Brain D, Carson D, Cauffman S, Chamberlin P, Chaufray J Y, Cheatom O, Clarke J, Connerney J, Cravens T, Curtis D, Delory G, Demcak S, DeWolfe A, Eparvier F, Ergun R, Eriksson A, Espley J, Fang X, Folta D, Fox J, Gomez-Rosa C, Habenicht S, Halekas J, Holsclaw G, Houghton M, Howard R, Jarosz M, Jedrich N, Johnson M, Kasprzak W, Kelley M, King T, Lankton M, Larson D, Leblanc F, Lefevre F, Lillis R, Mahaffy P, Mazelle C, McClintock W, McFadden J, Mitchell D L, Montmessin F, Morrissey J, Peterson W, Possel W, Sauvaud J A, Schneider N, Sidney W, Sparacino S, Stewart A I F, Tolson R, Toublanc D, Waters C, Woods T, Yelle R, Zurek R. 2015. The Mars atmosphere and volatile evolution (MAVEN) mission. Space Sci Rev, 195: 3–48

Jiang C, Yokoyama T, Wei L, Yang G, Zhao Z. 2021. Nonlinear simulation of ionospheric irregularities at Mars. Astrophys J, 909: 47

Keskinen M J. 2018. New model for ionospheric irregularities at Mars. Geophys Res Lett, 45: 2177–2183

Kil H, Paxton L J, Jee G, Nikoukar R. 2019. Plasma blobs associated with medium-scale traveling ionospheric disturbances. Geophys Res Lett, 46: 3575–3581

Kim V P, Hegai V V. 2016. Low latitude plasma blobs: A review. J Astron Space Sci, 33: 13–19

Krall J, Huba J D, Joyce G, Yokoyama T. 2010. Density enhancements associated with equatorial spread F. Ann Geophys, 28: 327–337

Krymskii A M, Breus T K, Ness N F, Hinson D P, Bojkov D I. 2003. Effect of crustal magnetic fields on the near terminator ionosphere at Mars: Comparison of in situ magnetic field measurements with the data of radio science experiments on board Mars Global Surveyor. J Geophys Res, 108: 1431

Laundal K M, Hatch S M, Moretto T. 2019. Magnetic effects of plasma pressure gradients in the upper F region. Geophys Res Lett, 46: 2355–2363

Le G, Huang C S, Pfaff R F, Su S Y, Yeh H C, Heelis R A, Rich F J, Hairston M. 2003. Plasma density enhancements associated with equatorial spread F: ROCSAT-1 and DMSP observations. J Geophys Res, 108: 1318

Lillis R J, Fillingim M O, Ma Y, Gonzalez-Galindo F, Forget F, Johnson C L, Mittelholz A, Russell C T, Andersson L, Fowler C M. 2019. Modeling wind-driven ionospheric dynamo currents at Mars: Expectations for InSight magnetic field measurements. Geophys Res Lett, 46: 5083–5091

Lühr H, Rother M, Maus S, Mai W, Cooke D. 2003. The diamagnetic effect of the equatorial Appleton anomaly: Its characteristics and impact on geomagnetic field modeling. Geophys Res Lett, 30: 1906

Mahaffy P R, Benna M, King T, Harpold D N, Arvey R, Barciniak M, Bendt M, Carrigan D, Errigo T, Holmes V, Johnson C S, Kellogg J, Kimvilakani P, Lefavor M, Hengemihle J, Jaeger F, Lyness E, Maurer J, Melak A, Noreiga F, Noriega M, Patel K, Prats B, Raaen E, Tan F, Weidner E, Gundersen C, Battel S, Block B P, Arnett K, Miller R, Cooper C, Edmonson C, Nolan J T. 2015. The neutral gas and ion mass spectrometer on the Mars atmosphere and volatile evolution mission. Space Sci Rev, 195: 49–73

Makela J J, Otsuka Y. 2012. Overview of nighttime ionospheric instabilities at low- and mid-latitudes: Coupling aspects resulting in structuring at the mesoscale. Space Sci Rev, 168: 419–440

Matta M, Mendillo M, Withers P, Morgan D. 2015. Interpreting Mars ionospheric anomalies over crustal magnetic field regions using a 2-D ionospheric model. J Geophys Res-Space Phys, 120: 766–777

Miller E S, Kil H, Makela J J, Heelis R A, Talaat E R, Gross A. 2014. Topside signature of medium-scale traveling ionospheric disturbances. Ann Geophys, 32: 959–965

Morschhauser A, Lesur V, Grott M. 2014. A spherical harmonic model of the lithospheric magnetic field of Mars. J Geophys Res-Planets, 119: 1162–1188

Němec F, Morgan D D, Gurnett D A, Duru F. 2010. Nightside ionosphere of Mars: Radar soundings by the Mars express spacecraft. J Geophys Res, 115: E12009

Němec F, Morgan D D, Gurnett D A, Brain D A. 2011. Areas of enhanced ionization in the deep nightside ionosphere of Mars. J Geophys Res, 116: E06006

Ness N F, Acuña M H, Connerney J E P, Kliore A J, Breus T K, Krymskii A M, Cloutier P, Bauer S J. 2000. Effects of magnetic anomalies discovered at Mars on the structure of the Martian ionosphere and solar wind interaction as follows from radio occultation experiments. J Geophys Res, 105: 15991–16004

Nielsen E, Fraenz M, Zou H, Wang J S, Gurnett D A, Kirchner D L, Morgan D D, Huff R, Safaeinili A, Plaut J J, Picardi G, Winningham J D, Frahm R A, Lundin R. 2007. Local plasma processes and enhanced electron densities in the lower ionosphere in magnetic cusp regions on Mars. Planet Space Sci, 55: 2164–2172

Niu D D, Cui J, Gu H, Wu X S, Wu S Q, Lu H Y, Chai L H, Wei Y. 2020. Energetic electron depletions in the nightside Martian upper atmosphere revisited. J Geophys Res-Space Phys, 125: e2019JA027670

Otsuka Y, Shiokawa K, Ogawa T, Wilkinson P. 2002. Geomagnetic conjugate observations of equatorial airglow depletions. Geophys Res Lett, 29: 43

Otsuka Y, Shiokawa K, Ogawa T, Wilkinson P. 2004. Geomagnetic conjugate observations of medium-scale traveling ionospheric disturbances at midlatitude using all-sky airglow imagers. Geophys Res Lett, 31: L15803

Park J. 2024. Climatology of Martian ionospheric disturbances deduced from MAVEN data in 2014–2023. J Geophys Res-Space Phys, 129: e2023JA031991

Park J, Min K W, Lee J J, Kil H, Kim V P, Kim H J, Lee E, Lee D Y. 2003. Plasma blob events observed by KOMPSAT-1 and DMSP F15 in the low latitude nighttime upper ionosphere. Geophys Res Lett, 30: 2114

Park J, Stolle C, Lühr H, Rother M, Su S Y, Min K W, Lee J J. 2008. Magnetic signatures and conjugate features of low-latitude plasma blobs as observed by the CHAMP satellite. J Geophys Res, 113: A09313

Park J, Lühr H, Stolle C, Rother M, Min K W, Chung J K, Kim Y H, Michaelis I, Noja M. 2009. Magnetic signatures of medium-scale traveling ionospheric disturbances as observed by CHAMP. J Geophys Res, 114: A03307

Penz T, Erkaev N V, Biernat H K, Lammer H, Amerstorfer U V, Gunell H, Kallio E, Barabash S, Orsini S, Milillo A, Baumjohann W. 2004. Ion loss on Mars caused by the Kelvin-Helmholtz instability. Planet Space Sci, 52: 1157–1167

Perkins F. 1973. Spread F and ionospheric currents. J Geophys Res, 78: 218–226

Picardi G, Biccari D, Seu R, Plaut J, Johnson W, Jordan R, Safaeinili A, Gurnett D, Huff R, Orosei R. 2004. MARSIS: Mars advanced radar for subsurface and ionospheric sounding. In: Wilson A, ed. Mars Express: A European Mission to the Red Planet. ESA Publ Div, Noordwijk. 51–69

Rodríguez-Zuluaga J, Stolle C, Yamazaki Y, Lühr H, Park J, Scherliess L, Chau J L. 2019. On the balance between plasma and magnetic pressure across equatorial plasma depletions. J Geophys Res-Space Phys, 124: 5936–5944

Schunk R, Nagy A. 2009. Ionospheres: Physics, Plasma Physics, and Chemistry. Cambridge: Cambridge University Press

Book   Google Scholar  

Siddle A G, Mueller-Wodarg I C F, Stone S W, Yelle R V. 2019. Global characteristics of gravity waves in the upper atmosphere of Mars as measured by MAVEN/NGIMS. Icarus, 333: 12–21

Soobiah Y, Coates A J, Linder D R, Kataria D O, Winningham J D, Frahm R A, Sharber J R, Scherrer J R, Barabash S, Lundin R, Holmström M, Andersson H, Yamauchi M, Grigoriev A, Kallio E, Koskinen H, Säles T, Riihelä P, Schmidt W, Kozyra J, Luhmann J, Roelof E, Williams D, Livi S, Curtis C C, Hsieh K C, Sandel B R, Grande M, Carter M, Sauvaud J A, Fedorov A, Thocaven J J, McKenna-Lawler S, Orsini S, Cerulli-Irelli R, Maggi M, Wurz P, Bochsler P, Krupp N, Woch J, Fränz M, Asamura K, Dierker C. 2006. Observations of magnetic anomaly signatures in Mars Express ASPERA-3 ELS data. Icarus, 182: 396–405

Stolle C, Lühr H, Rother M, Balasis G. 2006. Magnetic signatures of equatorial spread F as observed by the CHAMP satellite. J Geophys Res, 111: A02304

Sultan P J. 1996. Linear theory and modeling of the Rayleigh-Taylor instability leading to the occurrence of equatorial spread F. J Geophys Res, 101: 26875–26891

Tsunoda R T. 2006. On the coupling of layer instabilities in the nighttime midlatitude ionosphere. J Geophys Res, 111: A11304

Venkateswara Rao N, Mohanamanasa P, Rao S V B. 2017. Magnetically controlled density structures in the topside layer of the Martian ionosphere. J Geophys Res-Space Phys, 122: 5619–5629

Wan X, Xiong C, Wang H, Zhang K, Yin F. 2020. Spatial characteristics on the occurrence of the nighttime midlatitude medium-scale traveling ionospheric disturbance at topside ionosphere revealed by the Swarm satellite. J Geophys Res-Space Phys, 125: e2019JA027739

Wan X, Zhong J, Xiong C, Wang H, Liu Y, Li Q, Kuai J, Weng L, Cui J. 2021. Persistent occurrence of strip-like plasma density bulges at conjugate lower-mid latitudes during the September 8–9, 2017 geomagnetic storm. J Geophys Res-Space Phys, 126: e2020JA029020

Wan X, Zhong J, Zhang S R, Xiong C, Wang H, Liu Y, Huang F, Li Q, Kuai J, Chen J, Hao Y. 2022. Disturbance neutral winds effects on the ionospheric strip-like bulge at lower-middle latitudes. J Geophys Res-Space Phys, 127: e2022JA030541

Wan X, Zhong J, Wu Z, Hao Y, Xiong C, Wang H, He M, Cui J, Liu Y, Li Q, Kuai J. 2024a. Implication of kilometre-scale fluctuations on the MGS radio-occultation-derived electron density profiles: Artefacts or ionospheric wave-like structures. Mon Not R Astron Soc, 528: 4568–4581

Wan X, Zhong J, Hao Y, Cao Y, Cui J, Xiong C, Wang H, Liu Y, Kuai J, Li Q. 2024b. An automatic method for detection and naive classification of the Martian ionospheric irregularities. J Geophys Res-Space Phys, 129: e2023JA032228

Wang J S, Nielsen E. 2002. Possible hydrodynamic waves in the topside ionospheres of Mars and Venus. J Geophys Res, 107: 1039

Wang J S, Nielsen E. 2003. Wavelike structures in the Martian topside ionosphere observed by Mars Global Surveyor. J Geophys Res, 108: 5078

Wang X, Xu X, Ye Y, Wang J, Wang M, Zhou Z, Chang Q, Xu Q, Xu J, Luo L, He P, Cheng S. 2022. MAVEN observations of the Kelvin-Helmholtz instability developing at the ionopause of Mars. Geophys Res Lett, 49: e2022GL098673

Wang X, Xu X, Cui J, Gu H, Niu D, Zhou Z, Chang Q, Xu Q, Luo L, He P, Yi S. 2023. Electron density variability in the dayside ionosphere of Mars: The role of gravity waves. Mon Not R Astron Soc, 518: 4310–4321

Withers P, Fallows K, Girazian Z, Matta M, Häusler B, Hinson D, Tyler L, Morgan D, Pätzold M, Peter K, Tellmann S, Peralta J, Witasse O. 2012. A clear view of the multifaceted dayside ionosphere of Mars. Geophys Res Lett, 39: L18202

Xiong C, Yin F, Luo X, Jin Y, Wan X. 2019a. Plasma patches inside the polar cap and auroral oval: The impact on the spaceborne GPS receiver. J Space Weather Space Clim, 9: A25

Xiong C, Lühr H, Sun L, Luo W, Park J, Hong Y. 2019b. Long-lasting latitudinal four-peak structure in the nighttime ionosphere observed by the Swarm constellation. J Geophys Res-Space Phys, 124: 9335–9347

Xu S, Mitchell D, Liemohn M, Fang X, Ma Y, Luhmann J, Brain D, Steckiewicz M, Mazelle C, Connerney J, Jakosky B. 2017. Martian low-altitude magnetic topology deduced from MAVEN/SWEA observations. J Geophys Res-Space Phys, 122: 1831–1852

Download references

Acknowledgements

The MAVEN KP datasets are publicly archived in the Planetary Plasma Interactions (PPI) Node of NASA’s Planetary Data System (PDS) ( https://pds-ppi.igpp.ucla.edu/search/view/?f=yes&id=pds://PPI/maven.insitu.calibrated/data ). This work was supported by the Strategic Priority Research Program of Chinese Academy of Sciences (Grant No. XDB41000000), the National Natural Science Foundation of China (Grant Nos. 42374181, 42374186, 42104169, 42104147, 41804150, 41804153), the Guangdong Basic and Applied Basic Research Foundation (Grant Nos. 2021A1515011216, 2022A1515011580, 2020A1515110242), the Natural Science Foundation of Jiangsu Province (Grant No. BK20180445), the Joint Open Fund of Mengcheng National Geophysical Observatory (Grant No. MENGO-202217), the Science and Technology Project of Shangrao City (Grant No. 2021F002), the Key Laboratory of Tropical Atmosphere-Ocean System (Sun Yat-sen University), Ministry of Education, the Fundamental Research Funds for the Central Universities, the Opening Funding of Chinese Academy of Sciences dedicated for the Chinese Meridian Project, the Open Research Project of Large Research Infrastructures of CAS —“Study on the interaction between low/mid-latitude atmosphere and ionosphere based on the Chinese Meridian Project”.

Author information

Authors and affiliations.

Planetary Environmental and Astrobiological Research Laboratory (PEARL), School of Atmospheric Sciences, Sun Yat-sen University, Zhuhai, 519082, China

Xin Wan, Jiahao Zhong, Yongqiang Hao, Yutian Cao & Jun Cui

Key Laboratory of Tropical Atmosphere-Ocean System, Ministry of Education, Zhuhai, 519082, China

Xin Wan & Jiahao Zhong

Department of Space Physics, College of Electronic Information, Luojia Laboratory, Wuhan University, Wuhan, 430079, China

Chao Xiong & Hui Wang

School of Physics and Electronic Information, Shangrao Normal University, Shangrao, 334001, China

School of Petroleum, China University of Petroleum-Beijing at Karamay, Karamay, 834000, China

Qiaoling Li

College of Astronautics, Nanjing University of Aeronautics and Astronautics, Nanjing, 210016, China

Jiawei Kuai

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Jiahao Zhong .

Ethics declarations

Conflict of interest The authors declare that they have no conflict of interest.

Electronic Supplementary Material

Rights and permissions.

Reprints and permissions

About this article

Wan, X., Zhong, J., Hao, Y. et al. A general climatology of categorized Martian ionospheric irregularities. Sci. China Earth Sci. (2024). https://doi.org/10.1007/s11430-024-1388-x

Download citation

Received : 27 February 2024

Revised : 17 July 2024

Accepted : 19 July 2024

Published : 04 September 2024

DOI : https://doi.org/10.1007/s11430-024-1388-x

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Ionospheric irregularities
  • Crustal magnetic anomaly
  • Find a journal
  • Publish with us
  • Track your research

travelling ionospheric disturbances

ERAU researchers will launch 3 rockets during April 8 solar eclipse. Here's why.

T he April 8 total solar eclipse stretching from Texas to Maine provides Americans with more than a fleeting perceptible phenomenon.

It's also the setup for science.

Aroh Barjatya, a professor of engineering physics and director of the Space and Atmospheric Instrumentation Lab at Embry-Riddle Aeronautical University's Daytona Beach campus, and a team of researchers and students from ERAU and other institutions, will use the occasion to engage in a NASA-funded project called APEP, or Atmospheric Perturbations around Eclipse Path.

They will be working to gain a better understanding of the ionosphere ― part of the Earth's upper atmosphere in a belt that affects radio communications. The ionosphere starts at about 50 miles above the Earth's surface and extends beyond 350 miles, according to NOAA .

The ionosphere is beyond the stratosphere, where the ozone layer is located, as well as other layers. Because of the ionosphere's more direct exposure to the sun, particles from extreme ultraviolet and X-ray solar radiation, as well as cosmic rays, become charged, or ionized.

And because the ionosphere is so high above the Earth's surface, the pull of gravity is weakened, allowing the charged particles to be dynamic and swimming in plasma, the fourth state of matter, after solids, liquids and gasses. Barjatya compares the ionosphere to an ocean, as atmospheric conditions ― thunderstorms, hurricanes, tornadoes, even bombs exploding in Ukraine and Gaza ― can cause perturbations, or disturbances similar to oceanic waves.

"This layer reflects and refracts radio signals and impacts satellite communications as the signals pass through," Barjatya said.

So when people use GPS, perturbations in the ionosphere can affect the quality of the signal. Also, some Ham radio operators can use high-frequency signals to bounce off the ionosphere and reach listeners worldwide, so ionospheric disruptions can impact those communications, as well.

“All satellite communications and GPS-type navigation signals pass through the ionosphere and are affected in different ways depending on the conditions of the plasma there,” student-researcher Nathan Graves said in a university news release. "Thus, being able to understand ionospheric perturbations, model them and predict when they can happen is critically important in a connected society.”

Eclipses, too, can cause these perturbations. And for scientists like Barjatya, the time and location of eclipses are predictable, creating a perfect environment for targeted experiments.

Measuring solar eclipse's effect on ionosphere

In spring 2022, Barjatya and an Embry-Riddle research colleague, Shantanab Debchoudhury, were brainstorming ideas on how to take advantage of a NASA sounding rocket program, "Low Cost Access to Space," through which it launches about 20 rocket missions annually.

NASA had launched sounding rockets during eclipses seven previous times dating back 60 years, and two would occur within two years.

And ERAU's Space and Atmospheric Instrumentation Lab has experience launching sounding rockets and already had the design and one payload ready.

"We could design a neat experiment that involved launching three rockets in October 2023 in the annular eclipse, recovering and refurbishing the payload, and launching all three again in April 2024 during the total solar eclipse," Barjatya said.

Plus, Barjatya said: "The 2023 and 2024 eclipses were very close to existing sounding rocket ranges."

White Sands, New Mexico, and Wallops Island, Virginia, are two of NASA's launching sites. Last October, an annular solar eclipse was seen across New Mexico, where the launches were toward the eclipse path. The launches from Virginia's eastern seaboard will be away from the path, giving an additional perspective on how the ionopshere responds to the eclipse.

To collect more data from the ground and to model the findings, Barjatya found collaborators at Johns Hopkins University, Dartmouth College, the Massachusetts Institute of Technology and the Air Force Research Lab in New Mexico.

ERAU students built nine of the 10 instruments that will comprise the rockets' payloads. Dartmouth researchers contributed the other.

What will take place after liftoff

Here's how it will work: Most of the Embry-Riddle team will be at Wallops Island to launch three rockets, each 60 feet high with 18-inch diameters. One will lift off about 45 minutes before the peak local eclipse, the second will go off at peak eclipse and the third 45 minutes later.

The rockets are expected to soar to a height of 260 miles. There, they will open up to allow 10 instruments on board to collect data. They will also eject sub-payloads the size of a two-liter bottle of soda with instruments to collect data, allowing for the capture of 15 data sets at varying times during the eclipse.

The data gathered will measure plasma density, neutral density and magnetic fields. It will be transmitted to the ground using a system built and managed by NASA, Barjatya said.

Simultaneously, one of the ERAU doctoral students will be at White Sands to rerun ground experiments, while other ground-based teams from the Air Force Research Lab and the MIT Haystack Observatory, will gather larger-scale measurements from the ground, using various radar technologies.

Barjatya said the team will also release student-built, high-altitude balloons that will elevate to 19 miles above Earth, to help study lower atmosphere weather changes "and help us detangle how much of the ionosphere dynamics during eclipse could be from ground-based weather changes propagating upwards."

The ionosphere is not easy to study. Perturbations are unpredictable, but the eclipse is the exception.

"We know exactly when and where to launch the rockets to get some guaranteed scientific results. Most ionospheric missions must wait weeks to catch good science conditions to launch into, and sometimes the timing and positioning is still not perfect," Barjatya said. "I am excited by the precise launch timing and ability to comprehensively instrument a rocket platform."

Previous Rocket Launch: ERAU's Space and Atmospheric Instrumentation Lab has liftoff

Getting the Shot: Learn eclipse photography with USA Today Network photographers

Weather for Eclipse: Will it be cloudy during the solar eclipse in Florida? AccuWeather forecast says maybe

This article originally appeared on The Daytona Beach News-Journal: ERAU researchers will launch 3 rockets during April 8 solar eclipse. Here's why.

Embry-Riddle Aeronautical University scientists will join NASA in launching three sounding rockets from Wallops Island, Virginia, into the ionosphere during the April 8 solar eclipse. The rockets will be equipped with instruments seen here as they will be gathering data during the mission. From left are students Peter Ribbens and Nathan Graves and researchers Aroh Barjatya, the mission principal investigator and Robert Clayton, mission co-investigator.

IMAGES

  1. MAVEN Discovers Traveling Ionospheric Disturbances at Mars

    travelling ionospheric disturbances

  2. Traveling ionospheric disturbance signature in the electron isodensity

    travelling ionospheric disturbances

  3. 2018‐09‐04 traveling ionospheric disturbances (TIDs) observed with

    travelling ionospheric disturbances

  4. Two-dimensional maps of traveling ionospheric disturbances over North

    travelling ionospheric disturbances

  5. Schematic illustrating ionospheric response to natural and manmade

    travelling ionospheric disturbances

  6. Five other examples of traveling ionospheric disturbances at Mars

    travelling ionospheric disturbances

VIDEO

  1. Anumana Vs Nostromosis

  2. EL206 2. 3. 5 IONOSPHERIC ABSORPTION & DISTURBANCES KARTAR CH. GPC AJMER

  3. 12 Ionospheric data assimilation models, Ivan Galkin (BGD)

  4. Intercontinental travel will be cut off for a long period, Lonely Planet founder says

  5. Ionospheric irregularities & scintillations, Luca Spogli (INGV)

  6. RTN Best Practice pt2 at SELFPA-East in New Orleans on 8/15/12

COMMENTS

  1. Traveling Ionospheric Disturbances and Ionospheric Perturbations Associated With Solar Flares in September 2017

    Keograms of preflare and postflare traveling ionospheric disturbance propagation in ΔTEC (10 16 el/m 2 unit) in the zonal direction at fixed latitudes of ∼40° N, derived using 30-min sliding window filtering. Eastward propagation perpendicular to the direction of sunrise terminator (shown as the black line) can be identified during ...

  2. First Observations of Large Scale Traveling Ionospheric Disturbances

    Large Scale Traveling Ionospheric Disturbances (LSTIDs) are variations in the ionosphere with wavelengths greater than 1,000 km, periodicities between 30 min and 3 hr, and speeds greater than about 1,400 km per hour. Auroral zone disturbances are generally cited as the energy source for LSTIDs. In this paper, we show for the first time that ...

  3. Large‐Scale Traveling Ionospheric Disturbances Origin and Propagation

    We investigate the origin, occurrence, and propagation of large-scale traveling ionospheric disturbances (LSTIDs) over the European region during the 19-21 December 2015 geomagnetic storm. Analysis of the total electron content (TEC) perturbation component is supported by the ground-based Global Positioning System (GPS) and GLONASS (In ...

  4. Identification of Large-Scale Travelling Ionospheric Disturbances

    In this paper we analyze Digisonde observations obtained in the European region to specify the effects of large-scale travelling ionospheric disturbances (LSTIDs) on the ionospheric characteristics that define the conditions in the bottomside ionosphere. While this type of disturbances affects all frequency ranges in the F region, the most pronounced effect is detected in the foF2 critical ...

  5. Medium-scale traveling ionospheric disturbances by three-dimensional

    The traveling ionospheric disturbances (TIDs) are wave-like electron density disturbances that propagate through the ionosphere and cause wave-like TEC disturbances (Davies 1990; Kelley 2011) due to external energy input to the ionosphere and/or plasma instabilities. In mid-latitude, medium-scale TID (MSTID) appears frequently.

  6. PDF Large Scale Traveling Ionospheric Disturbances Observed using ...

    Large Scale Traveling Ionospheric Disturbances (LSTIDs) are quasi-periodic variations in F region electron density with horizontal wavelengths > 1000 km and periods between 30 to 180 min. On 3 November 2017, LSTID signatures were detected in simultaneously over the continental United

  7. A case study of the large-scale traveling ionospheric disturbances in

    Abstract. This study presents a comprehensive observation of the large-scale traveling ionospheric disturbances (LSTIDs) in the eastern Asian sector during the 2015 St. Patrick's Day (17 March 2015) geomagnetic storm. For the first time, three dense networks of GPS receivers in China and Japan are combined together to obtain the two-dimensional (2-D) vertical total electron content (VTEC ...

  8. Ionospheric Disturbance

    Ionospheric disturbances are a short-term deviation from regular climatology (such as diurnal variations). These disturbances can be local, regional, and sometimes global. A large category of these disturbances has characteristic wave properties in propagation and periodicity and is specifically termed as traveling ionospheric disturbances (TIDs).

  9. Statistical study of medium-scale traveling ionospheric disturbances in

    This study investigates the medium-scale traveling ionospheric disturbances (MSTIDs) statistically at the low-latitude equatorial ionization anomaly (EIA) region in the northern hemisphere. We apply the automatic detection algorithm including the three-dimensional fast Fourier transform (3-D FFT) and support vector machine (SVM) on total electron content (TEC) observations, derived from a ...

  10. IROTI: a new index to detect and identify traveling ionospheric

    The rate of total electron content index (ROTI), as the most common ionospheric disturbance index, is often used to detect equatorial plasma bubbles (EPBs). But it cannot effectively identify traveling ionospheric disturbances (TIDs). This research firstly proposes a new ionospheric disturbance index named the time integral and rate of the total electron content index (IROTI), which can be ...

  11. Sources and Measurement of Traveling Ionospheric Disturbances

    Traveling Ionospheric Disturbances (TIDs) are variations in the ionosphere that can impact medium frequency (MF) and high frequency (HF) radio communications through fading (QSB) and by causing variations in communications distance. HamSCI has two funded projects for studying TIDs and their connection to the atmosphere and space.

  12. Medium‐Scale Traveling Ionospheric Disturbances

    This paper provides a brief review of medium-scale traveling ionospheric disturbances (MSTIDs). MSTIDs have been thought to be caused exclusively by atmospheric gravity waves. However, novel imaging observations using optical and radio techniques revealed that MSTIDs during nighttime are not simply caused by the atmospheric gravity waves.

  13. GPS observations of medium-scale traveling ionospheric disturbances

    Abstract. Two-dimensional structures of medium-scale traveling ionospheric disturbances (MSTIDs) over Europe have been revealed, for the first time, by using maps of the total electron content (TEC) obtained from more than 800 GPS receivers of the European GPS receiver networks. From statistical analysis of the TEC maps obtained 2008, we have found that the observed MSTIDs can be categorized ...

  14. Solar activity dependence of medium-scale traveling ionospheric

    In order to reveal solar activity dependence of the medium-scale traveling ionospheric disturbances (MSTIDs) at midlatitudes, total electron content (TEC) data obtained from a Global Positioning System (GPS) receiver network in Japan during 22 years from 1998 to 2019 were analyzed. We have calculated the detrended TEC by subtracting the 1-h running average from the original TEC data for each ...

  15. Large-scale traveling ionospheric disturbances using ionospheric

    Study on large-scale traveling ionospheric disturbances (LSTIDs) triggered by solar activity in polar regions can improve the understanding of energy transmission and coupling from polar regions to mid- and low- latitude areas (Tsugawa et al., 2003). Previous statistic studies on ionospheric storms obtained general characters of the ...

  16. Observations of ionospheric irregularities during a geomagnetic storm

    This work was supported by the Special Found of Hubei Luojia Laboratory (220100011), the National Key R&D Program of China (Grant No. 2022YFF0503700), and the ISSI project "Resolving the Generation Mechanisms and Electrodynamical Effects of Medium Scale Traveling Ionospheric Disturbances (MSTIDs)".

  17. Short‐Period Concentric Traveling Ionospheric Disturbances Excited by

    Traveling ionospheric disturbances (TIDs) with different period scales that resulted from atmospheric acoustic gravity waves (AGWs) excited by rocket launches are frequently observed at ionospheric heights (Arendt, 1971; Li et al., 1994; Noble, 1990). Several types of TIDs can be observed during rocket launches.

  18. Travelling disturbances in the ionosphere

    Methods have been developed for the examination of the horizontal and vertical movements of short-period disturbances in the ionosphere. It has been found that quasi-periodic travelling disturbances with periods of from 10 to 60 min. are of frequent occurrence in the F region by day. They appear as temporary variations in the vertical distribution of ionization which show a horizontal ...

  19. Atmosphere

    In this study, we present the first results of detecting ionospheric irregularities using non-typical GPS observations recorded onboard the Geostationary Operational Environmental Satellites (GOES) mission operating at ~35,800 km altitude. Sitting above the GPS constellation, GOES can track GPS signals only from GPS transmitters on the opposite side of the Earth in a rather unique geometry ...

  20. A general climatology of categorized Martian ionospheric irregularities

    A climatological survey of Martian ionospheric plasma density irregularities was conducted by exploring the in-situ measurements of the Mars Atmosphere and Volatile EvolutioN (MAVEN) spacecraft. The irregularities were first classified as enhancement, depletion, and oscillation. By checking the simultaneous magnetic field fluctuation, the irregularities have been classified into two types ...

  21. Large‐Scale Traveling Atmospheric and Ionospheric Disturbances Observed

    Persistent LSTID/large-scale traveling atmospheric disturbances (LSTADs) are observed over several hours coinciding with moderate geomagnetic disturbances and a major sudden stratospheric warming (SSW) Along-track horizontal wavelengths of the atmospheric disturbances are ∼2,000-2,500 km and periods are ∼2-2.5 hr

  22. ERAU researchers will launch 3 rockets during April 8 solar ...

    Barjatya compares the ionosphere to an ocean, as atmospheric conditions ― thunderstorms, hurricanes, tornadoes, even bombs exploding in Ukraine and Gaza ― can cause perturbations, or ...

  23. Pilot Ionosonde Network for Identification of Traveling Ionospheric

    Traveling ionospheric disturbances (TIDs) are the ionospheric signatures of atmospheric gravity waves. Their identification and tracking is important because the TIDs affect all services that rely on predictable ionospheric radio wave propagation. Although various techniques have been proposed to measure TID characteristics, their real-time ...

  24. Numerical Study of Traveling Ionospheric Disturbances Generated by an

    Using a global atmosphere-ionosphere coupled model, the characteristics and excitation source of traveling ionospheric disturbances (TIDs) during geomagnetically quiet periods are studied. This is the first paper concerning the simulation of TIDs generated by upward propagating gravity waves (GWs) that are spontaneously generated in the model.