Skip to main content

DATA REPORT article

Front. Astron. Space Sci., 20 September 2023
Sec. Space Physics
Volume 10 - 2023 | https://doi.org/10.3389/fspas.2023.1240323

Interplanetary shock data base

  • 1Goddard Planetary Heliophysics Institute, University of Maryland, Baltimore County, Baltimore, MD, United States
  • 2Geospace Physics Laboratory, NASA Goddard Space Flight Center, Greenbelt, MD, United States

1 Introduction

Interplanetary (IP) shocks are frequently observed in the solar wind (Burlaga, 1971; Stone and Tsurutani, 1985). Many levels of different kinds of geomagnetic activity may follow the impact of IP shocks on the Earth’s magnetosphere. Such effects are seen everywhere in the magnetosphere-ionosphere system, including radiation belt dynamics, magnetic field in geosynchronous orbit, field-aligned currents, ionospheric disturbances, satellite orbital drag, ground magnetometers, geomagnetically induced currents (GICs), and others (e.g., Echer et al., 2005; Tsurutani et al., 2011; Khazanov, 2016; Oliveira and Ngwira, 2017; Oliveira and Zesta, 2019; Abda et al., 2020; Smith et al., 2020; Bhaskar et al., 2021). The study of IP shocks is important for space weather purposes because shock impacts occur more frequently than geomagnetic storms and correlate well with solar activity (Oh et al., 2007; Kilpua et al., 2015; Echer et al., 2023). Therefore, keeping an updated and accurate IP shock data base is of primary importance to the scientific community.

Many IP shock parameters control shock geoeffectiveness, such as shock speeds, Mach numbers, and compression ratios (Craven et al., 1986; Kabin, 2001; Goncharov et al., 2014). Additionally, the shock impact angle, the angle the shock normal vector performs with the Sun-Earth line, has been shown to be a significant factor that controls shock geoeffectiveness (Oliveira and Samsonov, 2018; Oliveira, 2023). Many works have demonstrated with simulations and observations that, in general, the more frontal and the faster the shock, the higher the subsequent geomagnetic activity observed from the geospace to the ground (e.g., Takeuchi et al., 2002; Guo et al., 2005; Wang et al., 2006; Oliveira and Raeder, 2014; Oliveira and Raeder, 2015; Samsonov et al., 2015; Oliveira et al., 2016; Selvakumaran et al., 2017; Oliveira et al., 2018; Baker, 2019; Rudd et al., 2019; Shi et al., 2019; Oliveira et al., 2020; Xu et al., 2020; Oliveira et al., 2021).

The main goal of this short report is to release an expanded version of an IP shock data base that was published before (Oliveira and Raeder, 2015; Oliveira et al., 2018). A major component of this new shock data base is a revision of the methodology used to calculate shock impact angles and speeds with respect to past versions of this list. Additionally, more shock and solar wind parameters before and after shock impacts and geomagnetic activity information were included in the list.

This article is organized as follows. Section 2 discusses the methodology used for the computation of shock properties, including the data used and shock normal calculation methods. A shock example is shown in Section 3. Section 4 presents the IP shock data base and its components. Finally, Section 5 brings a few suggestions for future usage of this shock list, with focus on the role of shock impact angles in controlling the subsequent shock geoeffectiveness.

2 Methods

2.1 Solar wind plasma and interplanetary magnetic field data

Properties and normal vector orientations of IP shocks are computed with the use of solar wind plasma and interplanetary magnetic field (IMF) data collected by solar wind monitors upstream of the Earth at the Lagrangian point L1. The time coverage of this shock list ranges from January 1995 to May 2023. Wind plasma data are collected by the Solar Wind Experiment instrument with resolution of 92 s (Ogilvie et al., 1995), and Wind magnetic field data are collected by the Magnetic Field Investigation instrument with resolution of 3 s (Lepping et al., 1995). ACE (Advanced Composition Explorer) collects solar wind data (resolution 64 s) with the Solar Wind Electron, Proton and Alpha Monitor instrument (McComas et al., 1998), and magnetic field data (resolution 16 s) with the MAG magnetometer instrument (Smith et al., 1998). All the data used for computations is represented in geocentric solar ecliptic (GSE) coordinates.

2.2 SuperMAG ground magnetometer and sunspot number data

Supporting geomagnetic index data are provided by the SuperMAG initiative (Gjerloev, 2009). SuperMAG computes geomagnetic indices using larger numbers of magnetometers in comparison to traditional IAGA (International Association of Geomagnetism and Aeronomy) indices (Davis and Sugiura, 1966; Rostoker, 1972). The SuperMAG ring current index, SMR, is explained by Newell and Gjerloev (2012), and the SuperMAG auroral indices SME, SMU, and SML are documented in Newell and Gjerloev (2011). SMU is the upper envelope index, SML is the lower envelope index, and SME = SMU—SML. All SuperMAG index data have resolution of 1 min.

Another supporting data set, with daily sunspot number observations, is provided by sunspot Index and Long-term Solar Observations, Royal Observatory of Belgium, Brussels (Clette and Lefèvre, 2016). The sunspot number data base used in this report has been corrected and recalibrated according to the methods explained by Clette and Lefèvre (2016).

2.3 Computations of shock impact angles and shock-related parameters

For the purpose of shock normal computations at 1 AU, shock fronts are assumed to be planar structures larger than the Earth’s magnetospheric system (Russell et al., 1983; Russell, 2000). Then, IP shock normal vectors can be computed if data from at least one spacecraft is available (Russell et al., 1983; Aguilar-Rodriguez et al., 2010; Trotta et al., 2023). Generally, shocks driven by CMEs (coronal mass ejections) have their shock normals with small deviation with respect to the Sun-Earth line, whereas shocks driven by CIRs (corotating interacting regions) have their shock normals with large deviations from the Sun-Earth line (Kilpua et al., 2015; Oliveira and Samsonov, 2018). Such shock inclinations occur because CMEs tend to travel radially in the solar wind, while CIRs tend to follow the Parker spiral when slow speed streams are compressed by fast speed streams (Pizzo, 1991; Tsurutani et al., 2006; Cameron et al., 2019). An animation showing the different inclinations of a CME-driven shock and a CIR-driven shock can be accessed here: https://dennyoliveira.weebly.com/phd.html.

There are three different ways commonly used to compute shock normal orientations. They use magnetic field data only, solar wind velocity data only, and a combination of magnetic field and solar wind velocity data. Such methods are, respectively, named magnetic coplanarity (MC, Colburn and Sonett, 1966), velocity coplanarity (VC, Abraham-Shrauner, 1972), and three mixed data methods (MX1, MX2, MX3, Schwartz, 1998). The equations used are listed below.

nMC=±B2×B1×B2B1B2×B1×B2B1(1)
nMX1=±B1×V2V1×B2B1B1×V2V1×B2B1(2)
nMX2=±B2×V2V1×B2B1B2×V2V1×B2B1(3)
nMX3=±B2B1×V2V1×B2B1B2B1×V2V1×B2B1(4)
nVC=±V2V1V2V1(5)

In these equations, B is the magnetic field vector, and V is the solar wind velocity vector. Indices 1 and 2 represent the upstream (non-shocked) region, and downstream (shocked) region, behind of and ahead the shock, respectively. The sign of each vector n is arbitrary and can be chosen to indicate whether the normal vector points toward the downstrean direction (+) or upstream direction (−) (Schwartz, 1998).

Eqs 15 provide a three-dimensional normal vector n=(nx,ny,nz) in Cartesian coordinates, from which three angles can be extracted.

θxn=cos1nx,(6)
φyn=tan1nzny,(7)
θBn=nB1|B1|,(8)

where θxn is named the shock impact angle, the angle the shock normal vector performs with the Sun-Earth line, φyn is the shock clock angle in the yz plane perpendicular to the Sun-Earth line (both angles in the satellite or Earth reference frame), and θBn is the angle between the upstream magnetic field vector and the shock normal vector (in the shock reference frame).

Since the plasma mass flux must be conserved along the shock normal, ρ1un1 = ρ2un2, with un1,n2=vsV1,2n, the shock speed is computed as follows:

vs=nV2ρ2V1ρ1ρ2ρ1(9)

Other useful shock velocities are represented by.

cs=γP1ρ1sound speed,(10)
vA=B1μ0ρ1Alfvén speed,(11)
vms=12vA2+cs2±vA2+cs224vA2cs2cos2θBnmagnetosonic speed,(12)

where γ = 5/3 is the ratio of the solar wind heat capacity with constant pressure to the heat capacity with constant volume; P1 is the upstream solar wind thermal pressure; ρ1 is the upstream solar wind density; and μ0 = 4π × 10–7 N/A2 is the magnetic vacuum permeability. The positive solution of Eq. 12 gives the fast magnetosonic speed, whereas the negative solution yields the slow magnetosonic speed (Jeffrey and Taniuti, 1964; Priest, 1981; Boyd and Sanderson, 2003).

Shock strengths are usually represented by specific Mach numbers. With u = vsVn being the relative speed between the shock speed and the local solar wind velocity, the Mach numbers are represented by.

MA=uvAAlfvénic Mach number,(13)
Ms=uvmsffast magnetosonic Mach number,(14)

where vmsf is the fast magnetosonic speed.

Finally, the strength of IP shocks can also be indicated by upstream and downstream solar wind plasma parameters and IMF. Such compression ratios are represented by.

Xn=n2n1plasma number density compression ratio,(15)
Xdp=ρ2V22ρ1V12dynamic pressure compression ratio,(16)
XB=|B2||B1|magnetic field compression ratio.(17)

3 The IP shock of 23 June 2000 as an example

Figure 1, first published by Oliveira and Raeder (2015), shows an IP shock event that occurred on 23 June 2000 observed by ACE at 1226 UT upstream of the Earth at (x, y, z) = (239.9, 36.7, −0.7) RE, where RE is the Earth’s radius = 6,371.1 km. Solar wind plasma and IMF data are depicted in the figure, along with SuperMAG geomagnetic index data. From top to bottom, the plot shows three components of the IMF (a); IMF magnitude (b); x component of solar wind velocity (c), y and z components of solar wind velocity (d); solar wind velocity magnitude (e); solar wind particle number density (f); solar wind dynamic pressure Pdyn = ρV2 (g); solar wind thermal temperature (h); SMR index (i); and SMU/SML indices (j). The vertical dashed magenta lines indicate the time of shock impact on the magnetosphere. The data were shifted to the magnetopause nose to match shock observations with the onsets in the ground geomagnetic indices. The highlighted grey areas correspond to the shock upstream region (left) 10 to 5 min before shock impact, and shock downstream region (right), 5 to 10 min after shock impact. Average values of these regions are used in Eqs 15 for the computation of the shock normal orientations with the five different methods. Most shocks have the time windows mentioned above, but a few events have different time windows.

FIGURE 1
www.frontiersin.org

FIGURE 1. Interplanetary shock observed by ACE on 23 June 2000 and the subsequent geomagnetic activity represented by SuperMAG data. (A) three components of the IMF; (B) IMF magnitude; (C) x component of solar wind velocity; (D) y and z components of solar wind velocity; (E) solar wind velocity magnitude; (F) solar wind particle number density; (G) solar wind ram pressure; (H) solar wind thermal plasma density; (I) SMR index; and (J) SMU and SML indices.

As discussed by Balogh et al. (1995) and Trotta et al. (2022), the length of the upstream and downstream windows around the shocks is an important factor in determining shock parameters. For example, Trotta et al. (2022) suggested a method to use windows with different lengths with short-length windows being located near the shock. This methodology provides statistical significance (including uncertainties) to the calculated shock parameters and reliability to the subsequent results. This approach will be applied to this shock list in a future work for further improvements of this shock data base.

The data shown in the figure is processed before plotting and before computing shock parameters and normal orientations. First, bad data points, such as 1E + 31 are replaced by nan (“not a number”) values and subsequently linearly interpolated. Then, solar wind parameter data are interpolated, and IMF data are averaged to a uniform time cadence of 30 s. Differences between the non-interpolated and interpolated data are very small or nearly nonexistent around the shock onset. This process allows the time resolutions of both data sets to match to further perform computations that involve both data sets. The same technique was applied to all events in the shock data base.

Positive step-like enhancements are seen in all solar wind plasma parameters and IMF. This is a clear signature of a fast forward IP shock (Priest, 1981; Tsurutani et al., 2011; Oliveira, 2017). More information on the analysis of this event including shock normal orientations will be provided in the next section.

4 The IP shock data base

Previous versions of this current shock data base were published by Oliveira and Raeder (2015) and Oliveira et al. (2018). A few sources were used to compile these previous shock lists: a shock catalog provided by the Harvard-Smithsonian Center for Astrophysics and compiled by Dr. J. C. Kasper for Wind (http://www.cfa.harvard.edu/shocks/wi_data/) and ACE (http://www.cfa.harvard.edu/shocks/ac_master_data/); a shock list compiled by the ACE team (at http://www-ssg.sr.unh.edu/mag/ace/ACElists/obs_list.html#shocks); and another list published by Wang et al. (2010) with events from February 1998 to August 2008. New events were added by scanning solar wind and IMF data to detect shock events that satisfy the framework discussed in Section 2.3.

The IP shock data base consists of three files: 1) full_shock_list_2023.txt, a text file with 603 events; 2) full_shock_params.cdf, a cdf file with detailed information about each specific shock event; and 3) read_shock.py, a file that contains a short python routine to read information about a specific shock event. The SpacePy package (https://spacepy.github.io, Morley et al., 2011; Larsen et al., 2022) is required to extract shock information from the cdf file using read_shock.py, which is shown in Listing 1.

Listing 1. Python routine to extract information related to a specific shock event from the cdf file.

www.frontiersin.org

The input variable for the above routine is the shock number sn. The IP shock represented in Figure 1 is the event number 142 in the shock list. Therefore, read_shock.py can be run as shown in Listing 2.

Listing 2. . Example of how to run the Python routine read_shock.py shown in Listing 1 to extract information about a shock event in the list. The example shown in this listing is the event number 142, occurred on 23 June 2000 and observed by ACE (see Figure 1).

www.frontiersin.org

The results for each shock are used to compose the shock list in the full_shock_list_2023.txt file. The file brings a header with the names of the variables (Table 1). The list also includes the position (in RE) of the solar wind monitor (either Wind or ACE) whose data are used in the calculations, along with minimum SMR values occurring in a time window of 2 hours after shock impact. This time window was chosen because amplitudes of geomagnetic activity response usually occur ∼60 min after energy being released by the magnetotail (Bargatze et al., 1985; Oliveira and Raeder, 2015; Oliveira et al., 2021). Such values can be used in studies that aim to use shock observations during non-storm times.

Below are the steps taken to include a specific solution for each shock in the full_shock_list_2023.txt list. These are the major revisions made to the list in comparison to its previous versions.

1 A filter is passed on the data to replace bad data points (e.g., 1E+31) by nan values which are then replaced by interpolated/averaged values to a common time cadence for both data sets (IMF and solar wind parameter data).

2 The satellite (Wind or ACE) must be in the solar wind upstream of the Earth (x > 14 RE).

3 If data of both satellites are simultaneously available, the data set with data of superior quality is used for computations.

4 Events with either MA or Ms (or both) smaller than one are generally discarded. Events with such conditions are only included in the list if they trigger significant geomagnetic activity, such as SMR variations of at least 15 nT.

5 The solution obtained from Eqs 15 closest to the median value is selected to be included in the list. If the difference between the maximum and minimum values of θxn is larger than 30°, the solution chosen for the list will be the one that shows more agreement with ground geomagnetic response, represented by the SuperMAG indices (SMR, SMU, SML, SME), as shown in previous publications (e.g., Wang et al., 2006; Oliveira and Raeder, 2015; Rudd et al., 2019; Oliveira et al., 2021).

The list version published by Oliveira and Raeder (2015) had 461 events, whereas the list published by Oliveira et al. (2018) had 547 events. This current data base has more events (603) and has a number of additional solar wind parameters and shock properties, which are shown in Table 1. The list time span, January 1995 to May 2023, includes two entire solar cycles (SC23 and SC24), the end of declining phase of SC22, and the beginning of ascending phase of SC25. Therefore, this data base provides a solid number of events for future statistical studies given an appropriate availability of data sets to be investigated.

Figure 2 represents general statistical features of the 603 events in the shock data base. Panel a shows yearly shock number distributions and Carrington-rotation of 25.38 days (Carrington, 1863) averaged sunspot numbers (solid black line). This figure shows a clear correlation between the number of events and sunspot numbers, a result that is already well known (Kilpua et al., 2015; Oliveira and Raeder, 2015; Rudd et al., 2019). Observations show that SC23 was significantly stronger than SC24 based on the overall sunspot numbers, which is reflected on the total number of shocks observed in the corresponding periods (432 and 187, respectively). Zhu et al. (2022) predicted that SC25 will be stronger than SC24 and will reach its maximum value around July 2025. Therefore, according to these predictions, it is reasonable to expect that SC25 will have a similar number of shocks with respect to SC24.

The overall statistical properties of sunspot observations and shock parameters are quantified and shown in Table 2. These results, including mean, median, percentile values, and shock number distribution correlation with sunspot numbers, are in excellent agreement with past studies (Kilpua et al., 2015; Oliveira and Raeder, 2015; Oliveira et al., 2018; Rudd et al., 2019).

TABLE 1
www.frontiersin.org

TABLE 1. Names of the variables associated with each shock event in the list and shown in Listing 2. The numbers in the first column are the numbers of the fields in the list and shown in the header of the file full_shock_list_2023.txt.

FIGURE 2
www.frontiersin.org

FIGURE 2. Statistical properties of shocks in the data base from January 1995 to May 2023 (603 shocks). Panel (A) shock number distribution and Carrington rotation-averaged sunspot numbers. Panels (B–J): number distributions of shock parameters obtained from Eqs 617.

TABLE 2
www.frontiersin.org

TABLE 2. Upper part: Shock parameters calculated from the shock data base released with this publication. The statistical data are: LPT, lower percentile (20%); median value; mean value; UPT, upper percentile (80%); and standard deviation (STD). There are 603 events in this shock list. Lower part: statistical results of sunspot number observations of the two complete solar cycles (SC) covered in the shock data base: SC23 and SC24.

5 Suggested use of the IP shock data base

The IP shock data base described in this report can be used in many ways. For example, this list can be used in studies involving geomagnetic activity following shock impacts from the geospace to the ground, including particle acceleration by shocks, particle dynamics and energization in the radiation belts, magnetospheric ultra-low frequency (ULF) waves, magnetic field response in geosynchronous orbit, field-aligned currents, role of shocks in substorm triggering, ionospheric irregularities, high-latitude thermosphere response (neutral density and nitric oxide) to shock impacts, ground magnetometer response (dB/dt variations) and subsequent effects on GICs, and many others. Therefore, this shock list can be used in a variety of space physics and space weather investigations.

A major feature of this shock list is the possibility of using the shock impact angle as a factor controlling geomagnetic activity. Oliveira (2023) has recently reviewed the effects of shock impact angles on the subsequent geomagnetic activity and also suggested a few topics for future research.

1 Shock impact angle effects on intensities and latitudinal extensions of dB/dt variations linked to enhancements of GICs (Carter et al., 2015; Oliveira et al., 2018; Oliveira et al., 2021).

2 Role of shock inclinations in controlling the triggering and wave modes of ULF waves and their interaction with magnetospheric cold plasma and wave-particle interactions (Oliveira et al., 2020; Hartinger et al., 2022).

3 Effects caused by different shock orientations on thermospheric neutral and nitric oxide molecules that control thermosphere heating and cooling affecting the subsequent satellite orbital drag in low-Earth orbit (Oliveira and Zesta, 2019; Zesta and Oliveira, 2019).

4 Shock impact angle effects on the dynamics of radiation belts (e.g., particle acceleration, enhancements, dropouts, and loss of relativistic electrons in the magnetosphere) (Tsurutani et al., 2016; Hajra and Tsurutani, 2018b).

5 Role of shock impact angle in triggering magnetospheric super substorms, with minimum SML < −2,500 nT (Hajra and Tsurutani, 2018a; Tsurutani and Hajra, 2023).

Finally, I would like to urge researchers to perform numerical simulations of shocks with different orientations. For example, Welling et al. (2021) argued that the “most perfect” CME would be very fast and impact Earth head-on. They performed numerical simulations of the impact of a perfect CME on the magnetosphere and concluded that ground dB/dt variations were noted in very low latitude regions because the CME impact was purely frontal. Furthermore, our shock list can be very useful in simulations comparing real observations with results yielded by numerical simulations of shocks with different orientations for many different space weather purposes.

Data availability statement

The interplanetary shock data base is available at https://zenodo.org/record/7991430. The solar wind plasma data and interplanetary magnetic field data were downloaded from NASA’s Coordinated Data Analysis Web (CDAWeb) website (https://cdaweb.gsfc.nasa.gov). SuperMAG geomagnetic index data were downloaded from https://supermag.jhuapl.edu. Daily sunspot number data files were downloaded from the Solar Influences Data Analysis Center of the Royal Observatory of Belgium (https://www.sidc.be/SILSO/datafiles).

Author contributions

The author confirms being the sole contributor of this work and has approved it for publication.

Funding

This work was possible thanks to the financial support provided by the NASA HGIO program through grant 80NSSC22K0756.

Conflict of interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Abda, Z. M. K., Aziz, N. F. A., Kadir, M. Z. A. A., and Rhazali, Z. A. (2020). A review of geomagnetically induced current effects on electrical power system: principles and theory. IEEE Access 8, 200237–200258. doi:10.1109/ACCESS.2020.3034347

CrossRef Full Text | Google Scholar

Abraham-Shrauner, B. (1972). Determination of magnetohydrodynamic shock normals. J. Geophys. Res. 77, 736–739. doi:10.1029/JA077i004p00736

CrossRef Full Text | Google Scholar

Aguilar-Rodriguez, E., Blanco-Cano, X., Russell, C. T., Jian, L. K., Luhmann, J. G., and Velez, J. C. R. (2010). “Study of interplanetary shocks using multi-spacecraft observations,” in Twelth international solar wind conference, AIP conference proceedings. Editors M. Maksimovic, K. Issautier, N. Meyer-Vernet, M. Moncuquet, and F. Pantellini (Washington, D.C.: American Institute of Physics), 1216, 467–470. doi:10.1063/1.3395904

CrossRef Full Text | Google Scholar

Baker, A. B. (2019). Effect of interplanetary shock impact angle on the occurrence rate and properties of Pc5 waves observed by high-latitude ground magnetometers. Master’s thesis. Blacksburg, Virginia: Virginia Tech.

Google Scholar

Balogh, A., Gonzalez-Esparza, J. A., Forsyth, R. J., Burton, M. E., Goldstein, B. E., Smith, E. J., et al. (1995). Interplanetary shock waves: ulysses observations in and out of the ecliptic plane. Space Sci. Rev. 72, 171–180. doi:10.1007/BF00768774

CrossRef Full Text | Google Scholar

Bargatze, L. F., Baker, D. N., McPherron, R. L., and Hones, E. W. (1985). Magnetospheric impulse response for many levels of geomagnetic activity. J. Geophys. Res. 90, 6387–6394. doi:10.1029/JA090iA07p06387

CrossRef Full Text | Google Scholar

Bhaskar, A., Sibeck, D., Kanekal, S. G., Singer, H. J., Reeves, G., Oliveira, D. M., et al. (2021). Radiation belt response to fast reverse shock at geosynchronous orbit. Astrophysical J. 910, 154. doi:10.3847/1538-4357/abd702

CrossRef Full Text | Google Scholar

Boyd, T. J. M., and Sanderson, J. J. (2003). The physics of plasmas. Cambridge, United Kingdom: Cambridge University Press.

Google Scholar

Burlaga, L. F. (1971). Hydromagnetic waves and discontinuities in the solar wind. Space Sci. Rev. 12, 600–657. doi:10.1007/BF00173345

CrossRef Full Text | Google Scholar

Cameron, T. G., Jackel, B. J., and Oliveira, D. M. (2019). Using mutual information to determine geoeffectiveness of solar wind phase fronts with different front orientations. J. Geophys. Res. Space Phys. 124, 1582–1592. doi:10.1029/2018JA026080

CrossRef Full Text | Google Scholar

Carrington, R. C. (1863). Observations of the spots on the Sun, from november 9th 1853 to march 24th 1861 made at redhill. London, United Kingdom: Williams and Norgate.

Google Scholar

Carter, B. A., Yizengaw, E., Pradipta, R., Halford, A. J., Norman, R., and Zhang, K. (2015). Interplanetary shocks and the resulting geomagnetically induced currents at the equator. Geophys. Res. Lett. 42, 6554–6559. doi:10.1002/2015GL065060

CrossRef Full Text | Google Scholar

Clette, F., and Lefèvre, L. (2016). The new sunspot number: assembling all corrections. Sol. Phys. 291, 2629–2651. doi:10.1007/s11207-016-1014-y

CrossRef Full Text | Google Scholar

Colburn, D. S., and Sonett, C. P. (1966). Discontinuities in the solar wind. Space Sci. Rev. 5, 439–506. doi:10.1007/BF00240575

CrossRef Full Text | Google Scholar

Craven, J. D., Frank, L. A., Russell, C. T., Smith, E. E., and Lepping, R. P. (1986). “Global auroral responses to magnetospheric compressions by shocks in the solar wind: two case studies,” in Solar wind-magnetosphere coupling. Editors Y. Kamide, and J. A. Slavin (Tokyo, Japan: Terra Scientific), 367–380.

CrossRef Full Text | Google Scholar

Davis, T. N., and Sugiura, M. (1966). Auroral electrojet activity index AE and its universal time variations. J. Geophys. Res. 71, 785–801. doi:10.1029/JZ071i003p00785

CrossRef Full Text | Google Scholar

Echer, E., de Lucas, A., Hajra, R., de Souza Franco, A. M., Bolzan, M. J. A., and do Nascimento, L. E. S. (2023). Geomagnetic activity following interplanetary shocks in solar cycles 23 and 24. Braz. J. Phys. 53, 1–13. doi:10.1007/s13538-023-01294-w

CrossRef Full Text | Google Scholar

Echer, E., Gonzalez, W. D., Guarnieri, F. L., Dal Lago, A., and Vieira, L. E. A. (2005). Introduction to space weather. Adv. Space Res. 35, 855–865. doi:10.1016/j.asr.2005.02.098

CrossRef Full Text | Google Scholar

Gjerloev, J. W. (2009). A global ground-based magnetometer initiative. Eos Trans. AGU 90, 230–231. doi:10.1029/2009EO270002

CrossRef Full Text | Google Scholar

Goncharov, O., Safrankova, J., Nemecek, Z., Prech, L., Pitna, A., and Zastenker, G. N. (2014). Upstream and downstream wave packets associated with low-Mach number interplanetary shocks. Geophys. Res. Lett. 41, 8100–8106. doi:10.1002/2014GL062149

CrossRef Full Text | Google Scholar

Guo, X.-C., Hu, Y.-Q., and Wang, C. (2005). Earth’s magnetosphere impinged by interplanetary shocks of different orientations. Chin. Phys. Lett. 22, 3221–3224. doi:10.1088/0256-307X/22/12/067

CrossRef Full Text | Google Scholar

Hajra, R., and Tsurutani, B. T. (2018a). Interplanetary shocks inducing magnetospheric supersubstorms (SML<-2500 nT): unusual auroral morphologies and energy flow. Astrophysical J. 858, 123. doi:10.3847/1538-4357/aabaed

CrossRef Full Text | Google Scholar

Hajra, R., and Tsurutani, B. T. (2018b). “Magnetospheric “killer” relativistic electron dropouts (REDs) and repopulation: A cyclical process,” in Extreme events in geospace: Origins, predictability and consequences. Editor N. Buzulukova (Cambridge, MA: Elsevier), 373–400. doi:10.1016/B978-0-12-812700-1.00014-5

CrossRef Full Text | Google Scholar

Hartinger, M. D., Takahashi, K., Drozdov, A. Y., Shi, X., Usanova, M. E., and Kress, B. (2022). ULF wave modeling, effects, and applications: accomplishments, recent advances, and future. Front. Astronomy Space Sci. 9. doi:10.3389/fspas.2022.867394

CrossRef Full Text | Google Scholar

Jeffrey, A., and Taniuti, T. (1964). Nonlinear wave propagation. New York, NY: Academic Press.

Google Scholar

Kabin, K. (2001). A note on the compression ratio in MHD shocks. J. Plasma Phys. 66, 259–274. doi:10.1017/S0022377801001295

CrossRef Full Text | Google Scholar

Khazanov, G. V. (2016). Space weather fundamentals. Boca Raton, FL: CRC Press.

Google Scholar

Kilpua, E. K. J., Lumme, K., Andréeová, E., Isavnin, A., and Koskinen, H. E. J. (2015). Properties and drivers of fast interplanetary shocks near the orbit of the Earth (1995-2013). J. Geophys. Res. Space Phys. 120, 4112–4125. doi:10.1002/2015JA021138

CrossRef Full Text | Google Scholar

Larsen, B. A., Morley, S. K., Niehof, J. T., and Welling, D. T. (2022). SpacePy. doi:10.5281/zenodo.3252523

CrossRef Full Text | Google Scholar

Lepping, R. P., Acuña, M. H., Burlaga, L. F., Farrell, W. M., Slavin, J. A., Schatten, K. H., et al. (1995). The WIND magnetic field investigation. Space Sci. Rev. 71, 207–229. doi:10.1007/BF00751330

CrossRef Full Text | Google Scholar

McComas, D. J., Bame, S. J., Barker, P., Feldman, W. C., Phillips, J. L., Riley, P., et al. (1998). Solar wind electron Proton Alpha monitor (SWEPAM) for the advanced composition explorer. Space Sci. Rev. 86, 563–612. doi:10.1023/A:1005040232597

CrossRef Full Text | Google Scholar

Morley, S. K., Koller, J., Welling, D. T., Larsen, B. A., Henderson, M. G., and Niehof, J. T. (2011). “Spacepy - a Python-based library of tools for the space sciences,” in Proceedings of the 9th Python in science Conference (SciPy 2010), Austin, TX, June 28 - July 3, 2010, 67–72.

Google Scholar

Newell, P. T., and Gjerloev, J. W. (2012). SuperMAG-based partial ring current indices. J. Geophys. Res. 117, 1–15. doi:10.1029/2012JA017586

CrossRef Full Text | Google Scholar

Newell, P. T., and Gjerloev, J. W. (2011). Evaluation of SuperMAG auroral electrojet indices as indicators of substorms and auroral power. J. Geophys. Res. 116. doi:10.1029/2011JA016779

CrossRef Full Text | Google Scholar

Ogilvie, K. W., Chornay, D. J., Fritzenreiter, R. J., Hunsaker, F., Keller, J., Lobell, J., et al. (1995). SWE, a comprehensive plasma instrument for the WIND spacecraft. Space Sci. Rev. 71, 55–77. doi:10.1007/BF00751326

CrossRef Full Text | Google Scholar

Oh, S. Y., Yi, Y., and Kim, Y. H. (2007). Solar cycle variation of the interplanetary forward shock drivers observed at 1 AU. Sol. Phys. 245, 391–410. doi:10.1007/s11207-007-9042-2

CrossRef Full Text | Google Scholar

Oliveira, D. M., and Samsonov, A. A. (2018). Geoeffectiveness of interplanetary shocks controlled by impact angles: A review. Adv. Space Res. 61, 1–44. doi:10.1016/j.asr.2017.10.006

CrossRef Full Text | Google Scholar

Oliveira, D. M., Arel, D., Raeder, J., Zesta, E., Ngwira, C. M., Carter, B. A., et al. (2018). Geomagnetically induced currents caused by interplanetary shocks with different impact angles and speeds. Space weather. 16, 636–647. doi:10.1029/2018SW001880

CrossRef Full Text | Google Scholar

Oliveira, D. M. (2023). Geoeffectiveness of interplanetary shocks controlled by impact angles: past research, recent advancements, and future work. Front. Astronomy Space Sci. 10. doi:10.3389/fspas.2023.1179279

CrossRef Full Text | Google Scholar

Oliveira, D. M., Hartinger, M. D., Xu, Z., Zesta, E., Pilipenko, V. A., Giles, B. L., et al. (2020). Interplanetary shock impact angles control magnetospheric ULF wave activity: wave amplitude, frequency, and power spectra. Geophys. Res. Lett. 47, e2020GL090857. doi:10.1029/2020GL090857

CrossRef Full Text | Google Scholar

Oliveira, D. M. (2017). Magnetohydrodynamic shocks in the interplanetary space: A theoretical review. Braz. J. Phys. 47, 81–95. doi:10.1007/s13538-016-0472-x

CrossRef Full Text | Google Scholar

Oliveira, D. M., and Ngwira, C. M. (2017). Geomagnetically induced currents: principles. Braz. J. Phys. 47, 552–560. doi:10.1007/s13538-017-0523-y

CrossRef Full Text | Google Scholar

Oliveira, D. M., and Raeder, J. (2014). Impact angle control of interplanetary shock geoeffectiveness. J. Geophys. Res. Space Phys. 119, 8188–8201. doi:10.1002/2014JA020275

CrossRef Full Text | Google Scholar

Oliveira, D. M., and Raeder, J. (2015). Impact angle control of interplanetary shock geoeffectiveness: A statistical study. J. Geophys. Res. Space Phys. 120, 4313–4323. doi:10.1002/2015JA021147

CrossRef Full Text | Google Scholar

Oliveira, D. M., Raeder, J., Tsurutani, B. T., and Gjerloev, J. W. (2016). Effects of interplanetary shock inclinations on nightside auroral power intensity. Braz. J. Phys. 46, 97–104. doi:10.1007/s13538-015-0389-9

CrossRef Full Text | Google Scholar

Oliveira, D. M., Weygand, J. M., Zesta, E., Ngwira, C. M., Hartinger, M. D., Xu, Z., et al. (2021). Impact angle control of local intense dB/dt variations during shock-induced substorms. Space weather. 19, e2021SW002933. doi:10.1029/2021SW002933

CrossRef Full Text | Google Scholar

Oliveira, D. M., and Zesta, E. (2019). Satellite orbital drag during magnetic storms. Space weather. 17, 1510–1533. doi:10.1029/2019SW002287

CrossRef Full Text | Google Scholar

Pizzo, V. J. (1991). The evolution of corotating stream fronts near the ecliptic plane in the inner solar system: 2. Three-Dimensional tilted-dipole fronts. J. Geophys. Res. 96, 5405–5420. doi:10.1029/91JA00155

CrossRef Full Text | Google Scholar

Priest, E. F. (1981). Solar magnetohydrodynamics. Dordrecht, Holland: D. Reidel Publishing.

Google Scholar

Rostoker, G. (1972). Geomagnetic indices. Rev. Geophys. 10, 935–950. doi:10.1029/RG010i004p00935

CrossRef Full Text | Google Scholar

Rudd, J. T., Oliveira, D. M., Bhaskar, A., and Halford, A. J. (2019). How do interplanetary shock impact angles control the size of the geoeffective magnetosphere? Adv. Space Res. 63, 317–326. doi:10.1016/j.asr.2018.09.013

CrossRef Full Text | Google Scholar

Russell, C. T., Gosling, J. T., Zwickl, R. D., and Smith, E. J. (1983). Multiple spacecraft observations of interplanetary shocks: ISEE three-dimensional plasma measurements. J. Geophys. Res. 88, 9941–9947. doi:10.1029/JA088iA12p09941

CrossRef Full Text | Google Scholar

Russell, C. (2000). The solar wind interaction with the Earth’s magnetosphere: A tutorial. IEEE Trans. Plasma Sci. 28, 1818–1830. doi:10.1109/27.902211

CrossRef Full Text | Google Scholar

Samsonov, A. A., Sergeev, V. A., Kuznetsova, M. M., and Sibeck, D. G. (2015). Asymmetric magnetospheric compressions and expansions in response to impact of inclined interplanetary shock. Geophys. Res. Lett. 42, 4716–4722. doi:10.1002/2015GL064294

CrossRef Full Text | Google Scholar

Schwartz, S. J. (1998). “Shock and discontinuity normals, Mach numbers, and related parameters,” in Analysis methods for multi-spacecraft data. Editors G. Paschmann, and P. W. Daly (Noordwijk, Netherlands: ESA Publications Division), 249–270. no. SR-001 in ISSI Scientific Report.

Google Scholar

Selvakumaran, R., Veenadhari, B., Ebihara, Y., Kumar, S., and Prasad, D. S. V. V. D. (2017). The role of interplanetary shock orientation on SC/SI rise time and geoeffectiveness. Adv. Space Res. 59, 1425–1434. doi:10.1016/j.asr.2016.12.010

CrossRef Full Text | Google Scholar

Shi, Y., Oliveira, D. M., Knipp, D. J., Zesta, E., Matsuo, T., and Anderson, B. (2019). Effects of nearly frontal and highly inclined interplanetary shocks on high-latitude field-aligned currents (FACs). Space weather. 17, 1659–1673. doi:10.1029/2019SW002367

CrossRef Full Text | Google Scholar

Smith, A. W., Rae, J., Forsyth, C., Oliveira, D. M., Freeman, P. M., and Jackson, D. (2020). Probabilistic forecasts of storm sudden commencements from interplanetary shocks using machine learning. Space weather. 18, e2020SW002603. doi:10.1029/2020SW002603

CrossRef Full Text | Google Scholar

Smith, C. W., L’Heureux, J., Ness, N. F., Acuña, M. H., Burlaga, L. F., and Scheifele, J. (1998). The ACE magnetic fields experiment. Space Sci. Rev. 86, 613–632. doi:10.1023/A:1005092216668

CrossRef Full Text | Google Scholar

R. G. Stone, and B. T. Tsurutani (Editors) (1985). “Collisionless shocks in the heliosphere: A tutorial review,” Geophysical monograph series (Washington, D.C: American Geophysical Union), 34. doi:10.1029/GM034

CrossRef Full Text | Google Scholar

Takeuchi, T., Russell, C. T., and Araki, T. (2002). Effect of the orientation of interplanetary shock on the geomagnetic sudden commencement. J. Geophys. Res. 107, SMP 6-1–SMP 6-10. SMP 6–1–SMP 6–10. doi:10.1029/2002JA009597

CrossRef Full Text | Google Scholar

Trotta, D., Hietala, H., Horbury, T., Dresing, N., Vainio, R., Wilson, L., et al. (2023). Multi-spacecraft observations of shocklets at an interplanetary shock. Mon. Notices R. Astronomical Soc. 520, 437–445. doi:10.1093/mnras/stad104

CrossRef Full Text | Google Scholar

Trotta, D., Vuorinen, L., Hietala, H., Horbury, T., Dresing, N., Gieseler, J., et al. (2022). Single-spacecraft techniques for shock parameters estimation: A systematic approach. Front. Astronomy Space Sci. 9. doi:10.3389/fspas.2022.1005672

CrossRef Full Text | Google Scholar

Tsurutani, B. T., Gonzalez, W. D., Gonzalez, A. L. C., Guarnieri, F. L., Gopalswamy, N., Grande, M., et al. (2006). Corotating solar wind streams and recurrent geomagnetic activity: A review. J. Geophys. Res. 111, A07S01–25. doi:10.1029/2005JA011273

CrossRef Full Text | Google Scholar

Tsurutani, B. T., and Hajra, R. (2023). Energetics of shock-triggered supersubstorms (SML <–2500 nT). Astrophysical J. 946, 17. doi:10.3847/1538-4357/acb143

CrossRef Full Text | Google Scholar

Tsurutani, B. T., Hajra, R., Tanimori, T., Takada, A., Remya, B., Mannucci, A. J., et al. (2016). Heliospheric plasma sheet (HPS) impingement onto the magnetosphere as a cause of relativistic electron dropouts (REDs) via coherent EMIC wave scattering with possible consequences for climate change mechanisms. J. Geophys. Res. Space Phys. 121 (10), 156. 130–10. doi:10.1002/2016JA022499

CrossRef Full Text | Google Scholar

Tsurutani, B. T., Lakhina, G. S., Verkhoglyadova, O. P., Gonzalez, W. D., Echer, E., and Guarnieri, F. L. (2011). A review of interplanetary discontinuities and their geomagnetic effects. J. Atmos. Solar-Terrestrial Phys. 73, 5–19. doi:10.1016/j.jastp.2010.04.001

CrossRef Full Text | Google Scholar

Wang, C., Li, C. X., Huang, Z. H., and Richardson, J. D. (2006). Effect of interplanetary shock strengths and orientations on storm sudden commencement rise times. Geophys. Res. Lett. 33 (14). doi:10.1029/2006GL025966

CrossRef Full Text | Google Scholar

Wang, C., Li, H., Richardson, J. D., and Kan, J. R. (2010). Interplanetary shock characteristics and associated geosynchronous magnetic field variations estimated from sudden impulses observed on the ground. J. Geophys. Res. 115. doi:10.1029/2009JA014833

CrossRef Full Text | Google Scholar

Welling, D. T., Love, J. J., Joshua Rigler, E., Oliveira, D. M., Komar, C. M., and Morley, S. K. (2021). Numerical simulations of the geospace response to a perfect interplanetary coronal mass ejection. Space weather. 19, e2020SW002489. doi:10.1029/2020SW002489

CrossRef Full Text | Google Scholar

Xu, Z., Hartinger, M. D., Oliveira, D. M., Coyle, S., Clauer, C. R., Weimer, D., et al. (2020). Inter-hemispheric asymmetries in the ground magnetic response to interplanetary shocks: the role of shock impact angle. Space weather. 18, e2019SW002427. doi:10.1029/2019SW002427

CrossRef Full Text | Google Scholar

Zesta, E., and Oliveira, D. M. (2019). Thermospheric heating and cooling times during geomagnetic storms, including extreme events. Geophys. Res. Lett. 46 (12), 12739–12746. 739–12. doi:10.1029/2019GL085120

CrossRef Full Text | Google Scholar

Zhu, H., Zhu, W., and He, M. (2022). Solar cycle 25 prediction using an optimized Long short-term memory mode with F10.7. Sol. Phys. 297, 157. doi:10.1007/s11207-022-02091-5

CrossRef Full Text | Google Scholar

Keywords: interplanetary shocks, data set, shock impact angle, solar cycle, shock parameters

Citation: Oliveira DM (2023) Interplanetary shock data base. Front. Astron. Space Sci. 10:1240323. doi: 10.3389/fspas.2023.1240323

Received: 14 June 2023; Accepted: 22 August 2023;
Published: 20 September 2023.

Edited by:

Steven Petrinec, Lockheed Martin Solar and Astrophysics Laboratory (LMSAL), United States

Reviewed by:

Fabio Lepreti, University of Calabria, Italy
Shuo Yao, China University of Geosciences, China

Copyright © 2023 Oliveira. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Denny M. Oliveira, denny@umbc.edu

Download