ABSTRACT

Little is known about the radio astronomical universe at frequencies below 10 MHz because such radiation does not penetrate the ionosphere. A Cubesat-based antenna for the 1–10 MHz band could be rapidly and economically deployed in low Earth orbit. When shielded by the Earth from Solar emission, it could observe weak extra-Solar System and outer Solar System sources. We suggest possibly observable sources, and application to study of the topside ionosphere.

1 INTRODUCTION

Low frequency (⁠|$\lesssim 25\,$| MHz) radio astronomy has been comparatively little studied because it is difficult to observe these frequencies from the ground. Science motivations for its study include very high redshift cosmology, low frequency sky surveys, solar/space weather, possible transient emissions from several classes of astronomical sources including fast radio bursts (FRB), soft gamma repeaters (SGR), gamma-ray bursts (GRB) and pulsars (PSR), and perhaps serendipitous discoveries.

The ionosphere reflects, at normal incidence, radiation at frequencies below a varying cutoff of 5–10 MHz, and at higher frequencies at grazing angles. Scintillation may be prohibitively strong even at frequencies at which the waves propagate. These values depend on the Solar cycle and activity, season and time of day, but largely preclude ground-based astronomy below 10 MHz. The lowest frequency successful ground-based observations appear to have been those of Bridle & Purton (1968) at 10.03 MHz, of Caswell (1976) at 10 MHz, of Cane (1979) at 5.2 MHz and of Ellis (1962) at 4.8 MHz, although some results at lower frequencies have been reported (Reber & Ellis 1956; Ellis 1957, 1965; Getmantsev et al. 1969). The abandonment of such low frequency ground-based observations more than 40 yr ago reflected a consensus that they are infeasible or not scientifically promising.

The ionosphere is not an obstacle to space-based observation. In fact, its existence is advantageous for observations at low frequencies because it shields space-based instruments from terrestrial electromagnetic interference while reflecting the sky. Bassett et al. (2020) observed extensive terrestrial interference above the ionosphere at frequencies 5–10 MHz, but very little at 3 MHz where the ionosphere is an effective shield. A few receivers of low frequency radiation have been flown: Alouette (Hartz 1964) in low Earth orbit (LEO), at a similar altitude to that proposed here but in a high-inclination rather than equatorial orbit, Radio Astronomy Explorer-1 (RAE-1) in a 5850 km orbit, higher than the LEO orbit contemplated here (Alexander et al. 1969; Weber, Alexander & Stone 1971) and Radio Astronomy Explorer-2 (RAE-2) in Lunar orbit (Alexander et al. 1975). The prospects of space-based low frequency radio astronomy were reviewed in a conference (Kassim & Weiler 1990). All observations of emission from outside the Solar System must contend with absorption and emission by interstellar plasma (Cong et al. 2021), and better characterization of this environment would be one of the scientific goals of such observations.

Several space instruments have observed Solar emissions at these low frequencies, including WAVES on the deep space Wind and STEREO spacecraft (Kaiser 2005; WAVES 2020; STEREO 2022). The Sun Radio Interferometer Space Experiment (SunRISE) (Kasper et al. 2019, 2022) is planned for geosynchronous orbit. Solar emission, including Type III bursts, might be a source of confusion or interference for observations of extra-Solar System sources. There can be no such confusion or interference when an antenna is shielded from the Sun by the Earth, which occurs for satellites in LEO with a duty cycle of about 40 per cent, but never, or rarely, in deep space or geosynchronous orbits (GEO).

After a long period of somnolence, technical and scientific progress argue for reviving space-based low frequency radio astronomy. The technical progress consists of the development of ‘Cubesats’, satellites consisting of one or more 10 cm cubes (Shkolnik 2018) and of microelectronics capable of sophisticated on-board data analysis within a small spatial envelope and with minimal power. Cubesats are simple enough that they are built as student projects. Launch into LEO may be free, piggybacking on other launches. The scientific progress includes the discovery (Lorimer et al. 2007) of FRB with durations |${\cal O} (\text{1 ms})$|⁠; the study of transients is a rapidly developing branch of radio astronomy. The behaviour of steady extra-Solar System radio sources at frequencies |$\lesssim 10\,$| MHz is also unknown; strong steady sources might be localized by a dipole antenna in LEO by occultation by the Earth’s limb, or by the Moon for favourably located sources (Andrew, Branson & Wills 1964).

Uncharacterized but spatially smooth deviations of the ionosphere from spherical symmetry might slightly limit the accuracy of source localization by this method, but would not greatly affect the strength of the reflected glint and of the peak of the autocorrelation. This peak would still distinguish a transient or rapidly varying source from the background of steady sources as well as from terrestrial interference, that may be significant even for low frequency exo-ionospheric observation of steady sources (Alexander et al. 1975).

Rajan et al. (2016) proposed a deep-space antenna array for low frequency radio astronomy, Sundkvist et al. (2016) proposed the CUbesat Radio Interferometry Experiment (CURIE) to observe Solar radio bursts, Yan, Wu, Gurvits et al. (2023) described the Low Frequency Interferometer and Spectrometer (LFIS) in Lunar orbit and Bentum et al. (2020) proposed the OLFAR (Orbiting Low Frequency Antennas for Radio Astronomy) system involving hundreds or thousands of satellites, linked to synthesize a large number of apertures. Burns (2020) proposed a Lunar-orbiting DAPPER and Lunar-surface based FARSIDE for low frequency observations, and Chen et al. (2020) proposed a low frequency interferometer array in Lunar orbit. Li et al. (2021) placed a lander, including a low frequency radio spectrometer, on the far side of the Moon and the necessary data relay satellite in orbit. These or similar projects offer the prospect of great scientific return some time in the future, but at high cost.

We propose a modest instrument that might, at less cost and sooner, perform a preliminary survey of |$\lesssim 10\,$| MHz radio astronomy. It would be based on a Cubesat in LEO with two centre-fed orthogonal half-wave dipole antennae at a nominal frequency of 3 MHz |$\lambda = 100\,$|m) these would be extended to lengths |$L = \lambda /4 = 25\,$| m by centrifugal force in each of four coplanar orthogonal directions. This nominal frequency is suggested because it is below the frequencies used by short-wave radio (Wikipedia 2022a) and likely also by over-the-horizon radar (Wikipedia 2022b); cf. the signals observed by RAE-2 in Lunar orbit (Alexander et al. 1975).

The orbital altitude is chosen above the peak of ionospheric electron density, where this density is low enough that it does not preclude transmission of extra-terrestrial radiation at the frequency of observation. A plasma frequency of 3 MHz (electron density |$n_e = 1.15 \times 10^5\,$| cm|$^{-3}$|⁠) in the topside ionosphere typically occurs at altitudes of 600–800 km (the peak electron density is several times higher and occurs at |$\approx$| 300–400 km) (International Reference Ionosphere 2022), depending on latitude, longitude, season, phase in the Solar cycle, time of day, and Solar activity. It is desirable to be above this critical altitude most of the time, so a nominal altitude |$h = 1000\,$| km is assumed. Refraction by the ionospheric plasma is significant at that altitude, but does not degrade the signal received by an antenna with little or no angular resolution, such as the dipole antennae considered.

This paper presents a theoretical analysis of the capabilities of the proposed instrument, based on the laws of physics, known properties of the ionosphere and possible but hypothetical extrapolated properties of astronomical radio sources. It does not contain the engineering design that would be required for a funding proposal. Most attention is paid to the instrument and its prospect of observing steady sources, such as supernova remnants, Galactic electron cosmic rays and active galactic nuclei. Interstellar and intergalactic dispersion and scatter broadening may preclude the detection of transients at frequencies |$\lesssim 10\,$| MHz, as discussed in an Appendix.

2 THE ANTENNA

A minimal system is shown in Fig. 1. Two orthogonal centre-fed half-wave (at the nominal frequency of 3 MHz) dipole antennae are extended from a Cubesat that contains amplifiers, data handling and storage electronics, and a higher frequency antenna for transmitting data to a ground station. Power is provided by Solar cells on the Cubesat. Larger telescopes with multiple half-wave antennae (separated by distances |$\sim \lambda /4$| to minimize capacitive coupling) could provide some angular resolution. The configuration is similar to that of Alouette (Hartz 1964), but the construction and deployment made compatible with a Cubesat.

Sketch of orbiting exoatmospheric radio telescope for observations at 1–10 MHz (numerical values in text apply for $\nu = 3\,$ MHz). Knudsen cell thrusters set the structure rotating, extending the wire antennae. The thrusters are aligned tangentially by ties to insulating diagonal aramid fibres, tightened by centrifugal force but that have no electromagnetic effects.
Figure 1.

Sketch of orbiting exoatmospheric radio telescope for observations at 1–10 MHz (numerical values in text apply for |$\nu = 3\,$| MHz). Knudsen cell thrusters set the structure rotating, extending the wire antennae. The thrusters are aligned tangentially by ties to insulating diagonal aramid fibres, tightened by centrifugal force but that have no electromagnetic effects.

2.1 Parameters

A minimum antenna wire radius is set by the requirement that resistive losses in the wire be small compared to its radiation resistance. For a resistance |$\Omega$| in an aluminum wire of length |$\lambda /2$| the wire radius

(1)

where the conductivity of aluminum |$\sigma _{\rm Al} = 3.6 \times 10^7\,$| mho m−1 (⁠|$\text{1 mho} = 1/\text{ohm}$|⁠). This corresponds to 26 AWG (American Wire Gauge). The mass of the two half-wave antennae is modest:

(2)

Past low frequency space antennae have been partial cylinders or tubes, extended, like a carpenter’s rule, by mechanical stiffness. This requires a comparatively thick, stiff, and massive antenna. The advantage of a centrifugally deployed wire antenna is that no mechanical thickness or stiffness is required; it is extended by tension. This permits a very thin and low mass antenna, limited only the requirement (equation 1) that it have sufficient electrical conductivity.

2.2 Sensitivity

The effective aperture of the proposed system for 3 MHz radiation polarized in its plane |$\sigma = \lambda ^2/4\pi \approx 800\,$| m|$^2$|⁠. This may be described by a gain |$G = \sigma /k_B \approx 0.6\,$| K Jy−1.1 The detection threshold is a flux density

(3)

where |$S/N$| is the required signal-to-noise ratio, |$\lambda = 100\,$| m the radio wavelength, |$T_{\rm rec}$|⁠, the noise temperature of the receiver, is scaled to the Galactic synchrotron sky brightness at 3 MHz (Alexander et al. 1969), |$G_{a}$| the telescope’s antenna gain (taken as unity for a dipole), B the receiving bandwidth, and |$t_{\rm int}$| the integration time.

2.3 Data download

A satellite in equatorial orbit passes over a near-equatorial ground station once per orbit. Data can be stored and downloaded with each passage over the ground station. The satellite would be within 2000 km for about 300 s each orbit, implying a minimum data transmission rate of |$\sim 5 \times 10^7$| samples per second during that overflight for a signal receiver bandwidth of 150 kHz (taken as 0.05 of the nominal observing frequency, in accord with the bandwidth of a dipole antenna). Assuming performance comparable to Planet Lab’s download rate of 280 Megabits s−1 would provide five bits per sample, or a dynamic range of |$2^5$|⁠.

The required mean power to transmit a dual polarization base-band signal sampled at a rate |$2 \pi \nu _{\rm obs}$| to a ground-based telescope of diameter D at a range R is

(4)

where |$S/N$| is the receiver signal-to-noise ratio, |$T_{\rm data}$| the data receiver noise temperature, and |$G_{\rm tr}$| is the transmitter gain. The ALMA Band 1 receiver, operating at 35–50 GHz, somewhat above the Ka (26–40 GHz) band, has a specified noise temperature of 32 K, justifying scaling to 30 K, but even a system noise temperature of 100 K would imply a power requirement of only |$\sim 0.2\,$| W.

The required power is modest, even with a dipole transmitting antenna (⁠|$G_{\rm tr} \approx 1$|⁠). Even a 1U Cubesat intercepts about 10 W of sunlight, and a larger satellite more. This could provide |$\gt 1\,$| W of photoelectric power with a duty factor (q.v. equation 17) |$\approx 0.67$|⁠.

The data transmission rate may be reduced by orders of magnitude if the received signal is processed on-board, taking advantage of Moore’s Law and the revolutions in electronics since the era of Alouette, RAE-1, and RAE-2. Rather than transmitting the base-band signal, it would only be necessary to transmit the scientific information of interest. This may be the total received power as a function of time, the autocorrelation of the received signal that indicates the presence of a transient event (because of the delay between direct and ionospheric-reflected signal), or the location of a steady source (again because the delay depends on its zenith angle).

3 ORBITAL LIFETIME

The mass of wire (equation 2) is small compared to the typical mass |$\sim n\,$|kg of a nU Cubesat, but the projected area of the antennae, each of length |$\lambda /2$|⁠, is |$2 \times 2 r \lambda /2 \approx 400\,$| cm|$^2$| (equation 1), several times the projected area of a Cubesat. Equating the work done by atmospheric drag to the decrease in energy of the satellite (noting that half the work done by gravity goes to increasing its kinetic energy), the orbital altitude h decreases at a rate

(5)

where |$R_{\rm orb}$| is the orbital radius (from the centre of the Earth), |$M_{{\rm Cube}}$| is the mass of the Cubesat, |$\rho _a$| is the atmospheric density, |$M_{{\oplus}} \approx 6.0 \times 10^{27}\,$| g is the mass of the Earth, and the drag coefficient |$C_d$| is taken as unity.

The scale height of the atmosphere at altitudes of interest (800–1200 km) is about 20 km because of its elevated temperature. As a result, the characteristic orbital lifetime

(6)

At these altitudes |$\rho _a$| is sensitive to the Solar cycle and activity (Jacchia 1970; Roberts 1971), and also depends on time of day (but not much on season at equatorial latitudes). It is more useful to specify the air density than the geometrical altitude, and it must be recognized that the orbital decay time (equation 6) may decrease rapidly and unpredictably with Solar activity.

4 DEPLOYMENT

The antennae must be extended by centrifugal force by setting the telescope rotating. Because of its small size, not much angular momentum can be imparted to the Cubesat by forces applied to its surfaces, but even a small initial angular momentum can begin the process of extension by rotating the Cubesat. As the antennae extend, thrusters at their ends produce increasing torques.

Several problems must be addressed:

  • The thrusters must be simple, light, and cheap.

  • The thrusters must continue to act over an extended time, perhaps hours or days, as the antennae gradually extend, increasing their lever arms.

  • The thrusters must remain tangentially oriented. The tiny torsional stiffness of the thin wire antennae that connect them to the Cubesat is insufficient to align them. Nor could tubular (or partial tubular) antennae be both stiff enough and have walls thick enough for handling within the mass budget.

The first two problems are solved by using Knudsen cells (Garland, Nibler & Shoemaker 2009) as the thrusters. A low vapour pressure compound, such as naphthalene, would gradually escape through an aperture at one end of each cell, with its recoil providing the thrust.

The third problem is solved by connecting the ends of the antennae with fine electrically insulating fibre, such as aramid, as shown in Fig. 1, and fixing the Knudsen cells to the fibres. Aramid fibres are available as thin as 170 dtex (1 dtex is defined as a mass of 1 g/10 km) corresponding to a radius of about |$60\, \mu = 0.006\,$| cm (this is also expressed as a length per unit mass |$\text{Nm} = 60$|⁠, where 1 Nm is 1 m g−1). These fibres have the negligible total mass of about 2.5 g. Each fibre has a tensile strength of tens of N, orders of magnitude greater than its tensile load at an angular rotation rate of 3.6 s−1 (equation 9); it is only necessary that the rotation rate be much greater than the orbital angular frequency |$\omega _{\rm orb} \approx 10^{-3}\,$| s|$^{-1}$| in low Earth orbit to maintain the geometry.

As the antennae extend these fibres will also be made taut by centrifugal force. Thrusters tied to them would be aligned tangentially, so their recoil forces spin up the entire system, keeping it taut and stable. The moment of inertia of the four-armed (two |$\lambda /2$| dipole antennae) telescope shown in Fig. 1 is

(7)

where equation (1) has been taken for the wire radius r.

Free molecular flow from Knudsen cells imparts an angular momentum

(8)

where |$m_p$| is the mass of propellant gas exhausted, |$m_g$| its molecular weight, and T its temperature T, and hence the vapour pressure and evaporation time, are determined by the radiative properties of the outsides of the Knudsen cells. For naphthalene at 300 K the rotation rate

(9)

For |$m_p = 10\,$| g the load on the wire at the Cubesat is |$\pi r^2 \omega \rho _{\rm Al} L^2/2 \approx 1.5\,$| N and the tensile stress |$\omega ^2 \rho _{\rm Al} L^2/2 \approx 1.1 \times 10^8\,$|dyne cm−2, less than a tenth of the tensile strength of aluminum. The peripheral velocity |$\omega L \approx 90 \,$| m s−1.

If the telescope plane is inclined at an angle i to its orbital plane its spin angular momentum and plane precess (as a result of the Earth’s gravitational torque) around its orbital angular momentum at a rate

(10)

or about one radian per month for the assumed parameters. Spin precession slews the broad dipole antenna pattern on the sky at an angular rate |$\omega _{\rm pre} \sin {i}$| with angular amplitude i. Significantly faster or slower precession can be obtained by choice of |$m_p$| and hence of the rotation rate |$\omega$|⁠. The rotational plane of a telescope whose orbit is not equatorial will also precess because of Earth’s equatorial bulge, but (if its spin and orbit are aligned) at a much slower rate than given by equation (10).

5 SOURCES

There are no known extra-Solar System point sources of radiation in the 1–10 MHz range, but sources observed at higher frequencies may also be detectable at these lower frequencies. Promising candidates include radio galaxies and AGN. The mean emission of the Crab pulsar was detected at 26.5 MHz with a flux density |$\sim 1000\,$| Jy by Andrew et al. (1964) before its discovery as a pulsar!

5.1 Attenuation

Galactic absorption is significant (Cane 1979) at frequencies of a few MHz, and at lower frequencies sets an effective horizon within the Galactic disc. This need not preclude observation of old neutron stars that may radiate at these lower frequencies, even though they are not detected in pulsar searches at VHF and UHF frequencies. There are |$\sim 10^8$| neutron stars in the Galactic disc, so that the nearest is likely at a distance of |$\approx 20\,$| pc. The path from such a close source has an absorption optical depth of only |$\sim 0.2$| of that through the full thickness of the Galactic disc, so the radiation from such a nearby source would be much less attenuated and scattered than that from outside the disc.

The opacity at 3 MHz (Spitzer 1962)

(11)

where the electron density |$n_e$| (in cm|$^{-3}$|⁠) is assumed to come from singly ionized species. In a weakly ionized (⁠|$n_e = 0.03\,$| cm|$^{-3}$|⁠) cool (100 K) cloud the absorption length is |$\sim 30\,$| pc and varies nearly |$\propto T^{3/2}$|⁠, while in a warm (⁠|$10^4$| K) ionized intercloud medium (⁠|$n_e = 0.01\,$| cm|$^{-3}$| in pressure equilibrium with a cool neutral cloud with |$n_\text{H} = 1\,$| cm|$^{-3}$|⁠) the absorption length is |$\sim 150\,$| kpc.

Cool clouds may be opaque, but cosmic ray electrons within them produce an internal source of radio radiation. Condensation into clouds likely increases their cosmic ray density and magnetic field, so they may be net emitters in comparison to the extra-Galactic background. The warm ionized intercloud medium is transparent, transmitting the extra-Galactic background.

5.2 Steady sources

Electron cosmic rays emit incoherent but non-thermal radio synchrotron radiation, and provide a well-understood background. The absorption of this radiation by interstellar ionized gas diagnoses the spatial and temperature distribution of that gas (Ellis & Hamilton 1966; Weber et al. 1971; Alexander et al. 1975; Cane 1979). The proposed system would extend these observations to lower frequencies where the interstellar plasma absorption, varying as the |$-2$| power of frequency, is greater.

Occultation by the Earth’s limb would permit angular resolution on small angular scales in one dimension, as in the early days of X-ray astronomy occultation by the Moon localized X-ray sources on the sky, and permitted their first identifications. This would enable mapping of the distribution of the Galactic non-thermal radio synchrotron radiation, and hence of the distribution of electron cosmic rays. The anisotropic response of the dipole antenna would permit low-resolution mapping.

It may also be possible to infer the absorption along paths to very bright discrete sources, such as Cygnus A, by observing the change of the sky-integrated signal as it enters or leaves Earth occultation. This radio galaxy has a synchrotron radiation spectrum with a flux density at 74 MHz of 1040 Jy and a spectral index between 74 and 327 MHz of −1.49 (Lazio et al. 2006). Even if there is a spectral turnover, a mean spectral index of −1 would imply a flux density of about 25 kJy. This steady source would be readily observable at 3 MHz. A 10 000 s integration would provide |$B t_{\rm int} \approx 1.5 \times 10^9$| and a detection threshold of about 1 kJy (equation 3). Its spectrum would be diagnostic of the interstellar medium along its line of sight.

An additional possible source might be interstellar magnetic reconnection, dissipating the energy of the interstellar field, amplified by magnetohydrodynamic turbulence. These processes are ubiquitous in conducting plasmas, from laboratory magnetic fusion reactors to the Solar photosphere and corona and wind, and stellar and quasar accretion discs. Electric fields resulting from changing magnetic fields accelerate electrons to relativistic energies, and plasma instabilities lead to turbulent resistivity. Electrons accelerated to Lorentz factors |${\cal O}(10^3)$| have synchrotron frequencies of about 3 MHz, making them possible sources of radiation in this band. It is unclear how rapidly such a source would vary; it might be steady on instrumental time-scales.

6 THE TOPSIDE IONOSPHERE

The spatial structure of the topside ionosphere (Banks, Schunk & Raitt 1976; Pignalberi et al. 2020; Prol et al. 2022) may be probed by observing the reflection of 3 MHz radiation from a strong steady natural source or an artificial beacon. Natural sources include the radio galaxies Cen A and Cyg A and the SNR Cas A. If the reflective layer is tilted by gravity waves (or otherwise), the source’s effective elevation varies, and can be inferred from the phase difference and interference between the reflected and direct signals from steady sources of known direction.

The low Galactic latitude Cas A (⁠|$b = -1.96^\circ$|⁠) may be observable only at frequencies |$\gtrapprox 10\,$| MHz because of interstellar absorption (Stanislavsky et al. 2023), that may also be significant for Cyg A (⁠|$b = 5.6^\circ$|⁠). At larger zenith angles (lower elevations) the ionosphere is reflective at higher frequencies, less absorbed by interstellar plasma, and grazing reflection at these frequencies can also be used to study the topside ionosphere.

In the simplest possible model the ionosphere is static (except for the effect of the diurnal variation of the Solar ionizing radiation) and its density contours are spherical (horizontal in the flat-ionosphere approximation). However, the elevation |$\Delta$| (and zenith angle |$\theta$| vary as the satellite moves in its orbit, so the path difference |$\delta \ell$| and phase difference |$2 \pi \delta \ell /\lambda$| vary with time, comparatively rapidly. For a steady source the received power will oscillate as the phase difference between these two paths changes as |$\Delta$| varies with the satellite motion. This is the same principle as that of the Sea Interferometer (Bolton & Slee 1953) of early radio astronomy, itself a long-wave realization of Lloyd’s mirror.

For an ideal static ionosphere the frequency of oscillation is determined by the satellite’s orbit and the direction to the source. Ionospheric oscillations (tilts of the reflecting surface, which is the critical density surface for sources at the zenith but is higher for sources at non-zero zenith angles) will manifest themselves as deviations of these oscillations from the predictions of the flat-ionosphere model. This would require only measurement of the received power, averaged over a fraction of the oscillation period (typically seconds), rather than processing or transmission of base-band data.

The elevation |$\Delta$| of a source at a known Right Ascension |$\alpha$| and Declination |$\delta$| as viewed from a point in an equatorial orbit (⁠|$\delta _{\rm orb} = 0$|⁠) with geocentric Right Ascension |$\alpha _{\rm orb} = \Omega t$|⁠, where |$\Omega$| is the orbital angular velocity, is given by

(12)

The amplitude of the sum of the direct and reflected fields, and hence of the received power, oscillates with a period |$P_{\rm osc}$|⁠. The rate of change of the delay between the two signals is

(13)

or

(14)

A full cycle of this oscillation occurs after a time

(15)

where |$\omega$| is the angular frequency of the radio radiation. This condition is met after an interval

(16)

where |$P_{\rm orb}$| is the orbital period. For |$P_{\rm orb} = 100\,$| min, |$\omega = 2 \times 10^7$| s|$^{-1}$| (3 MHz) and |$h = 300$| km, the numerical factor is about 0.3 s. As many as |${\cal O}(10^4)$| cycles of this oscillation could be observed in the approximately half-orbit (3000 s) during which an astronomical source can be observed, so that ionospheric tilts |${\cal O}(10^{-4})$| radian might be detectable.

The approximately 5 per cent bandwidth of a half-wave dipole is wide enough that the oscillation phase varies by many radians across the band, so that the band would have to be divided into channels to be analysed separately. The oscillation frequency (or |$P_{\rm osc}$|⁠) of the received power is described by few tens of bits of data because the oscillation has a frequency of a few Hz. Transmitting these stored data, even for hundreds of frequency channels, would not require much bandwidth.

The delay also requires correction for propagation through layers in which the electron density is high enough that the signal group velocity is significantly less than c, and its path is bent by refraction. Because the electron gyrofrequency is 1–2 MHz, the effect of the geomagnetic field is significant and the ordinary and extraordinary modes must be considered separately within the ionosphere.

This analysis depends on the assumption that the ionospheric density contours are smooth, undisturbed by turbulence. The failure of these predictions (variation of the received intensity that is not periodic with the period of equation 16) would therefore be a novel probe of topside ionospheric turbulence that is not addressed by existing ionospheric diagnostics (Schreiter et al. 2023).

7 SHIELDING OF THE SUN

At low frequencies the Sun is an intense source of background in a dipole antenna. The Earth shields this background whenever the antenna is in its shadow. At equinoxes it is shielded by the solid Earth a fraction

(17)

of the time, where we have taken |$h = 1000\,$| km.

The ionosphere increases |$f_{\rm shield}$| because it increases the effective (opaque) |$R_{{\oplus}}$|⁠. In addition, ionospheric refraction when the Sun is near the antenna’s horizon may further increase |$f_{\rm shield}$|⁠:

(18)

where

(19)

is the maximum angle of ionospheric plasma refraction, |$\nu _{p\text{-}\rm orb}$| is the plasma frequency at the antenna’s altitude and |$\nu$| is the radiation frequency. The depression angle of the critical density horizon

(20)

where |$h_{\rm crit}$| is the altitude at which the plasma frequency |$\nu _p(h_{\rm crit}) = \nu$|⁠; for observation to be possible |$h_{\rm crit} \lt h$|⁠.

Both |$\phi$| and |$\psi$| are small angles for a satellite in LEO. Averaging over the year, |$f_{\rm shield}$| is multiplied by a factor |$\approx 1-\varepsilon ^2/4 \approx 0.96$|⁠, where |$\varepsilon \approx 0.41\,$| rad is the obliquity of the Earth’s equator.

In contrast, deep space or GEO observatories like FIRST, SURO-LC, DARIS, Wind, STEREO, SunRISE, and OLFAR are illuminated by the Sun; their purpose is to observe it and its corona, but this background is a severe obstacle to observations of weak extra-Solar System sources. Refraction by interplanetary plasma that broadens the arrival directions of low frequency radiation by |$\Delta \theta \sim 10^\prime \sim 3\,$| mrad spreads its arrival time by |$\sim 1\, \text{AU} (\Delta \theta )^2/2c \sim 2.5\,$|ms, limiting the possible resolution of interferometry, even with long baselines.

8 DISCUSSION

Even without angular resolution, it would be possible to probe the Universe in this unexplored frequency band:

  • A dipole antenna would measure a sky average (weighted by its gain) temperature, yielding information about the interstellar medium not obtainable in any other manner. Models of the interstellar medium predict the antenna temperature of a dipole, and can be tested by its measurement.

  • A dipole antenna observing the reflection of an exospheric beacon would measure the variability of the topside ionosphere.

Angular resolution would provide additional information:

  • A single dipole telescope (or two orthogonal dipoles) whose rotation axis is not parallel to Earth’s would precess, even in equatorial orbit. This would sweep its dipole beam pattern across the sky, providing some angular resolution of steady emission like that of the interstellar medium.

  • Resolution could be obtained by aperture synthesis with a larger telescope comprising multiple dipoles. This could resolve interstellar cloud structure.

  • Aperture synthesis using multiple, widely spaced, dipoles in equatorial orbit could narrowly constrain the locations of any transients strong enough to be observed, but would be poorly matched to the broad angular scales of interstellar clouds. More closely spaced dipoles would be a better match to that angular structure. Knowledge of the dipoles’ locations (although not necessarily active station-keeping) to allow for unpredictable differences in atmospheric drag would be required; this could be provided by GPS.

An equatorial orbit would permit data downloads to a single ground station every orbit, minimizing data storage and download rate requirements. The orbit of such a satellite, locked to the Earth’s equatorial bulge, would not precess. If rotating in the Earth’s equatorial plane, its spin would also not precess.

Large, low frequency space-based antennae are plausible candidates for demonstration of in-space manufacturing and assembly. Such structures would need to be quite large (⁠|$\sim \,$|km) in order to provide even degree-scale resolution at these long wavelengths. Before such a costly and technically challenging demonstration it would be prudent to develop a ‘pathfinder’ Cubesat-scale telescope to explore the signal characteristics to be observed by a larger instrument.

An extended antenna wire poses a collision risk for other satellites. Its effective cross-section for a 1 m satellite would be |$\sim 50\,$| m|$^2$|⁠. This is much larger than the satellite’s |$\sim 1\,$| m|$^2$| cross-section for a small piece of space debris, but the proposal is only for one antenna while lethal debris are numerous. If the wire is thin enough collision might not be catastrophic to the satellite. Once the mission is over, the antennae could be detached from the Cubesat. No longer centrifugally extended, a slight (unavoidable) intrinsic curvature would crumple the thin wire into a compact tangle, with much reduced cross-section.

The mutual collison risk of multiple 3 MHz antennae, such as might be deployed to make an interferometer, would be minimized if all were in equatorial orbits, with rotation axes parallel to the Earth’s and each other’s. Then their mutual collision cross-section would be proportional only to the first power of their antenna length, rather than quadratic.

DATA AVAILABILITY

This theoretical work generated no original data.

ACKNOWLEDGEMENTS

We thank Joseph Lazio, David Palmer, and anonymous referees for calling our attention to earlier work on this subject.

Footnotes

1

This is a different meaning of ‘gain’ than that used by antenna engineers.

References

Alexander
 
J. K.
,
Brown
 
L. W.
,
Clark
 
T. A.
,
Stone
 
R. G.
,
Weber
 
R. R.
,
1969
,
ApJ
,
157
,
L163
 

Alexander
 
J.
,
Kaiser
 
M.
,
Novaco
 
J.
,
Grena
 
F.
,
Weber
 
R.
,
1975
,
A&A
,
40
,
365

Andrew
 
B. H.
,
Branson
 
N. J. B. A.
,
Wills
 
D.
,
1964
,
Nature
,
203
,
171
 

ATNF Pulsar Catalogue
,
2005
, Available at:

Banks
 
P. M.
,
Schunk
 
R. W.
,
Raitt
 
W. J.
,
1976
,
Ann. Rev. Earth Planet. Sci.
,
4
,
381
 

Bassett
 
N.
,
Rapetti
 
D.
,
Burns
 
J. O.
,
Tauscher
 
K.
,
MacDowall
 
R.
,
2020
,
Adv. Space Res.
,
66
,
1265
 

Bentum
 
M. J.
 et al. ,
2020
,
Adv. Space Res.
,
65
,
856
 

Bilous
 
A. V.
 et al. ,
2020
,
A&A
,
635
,
A75
 

Bochenek
 
C. D.
,
Ravi
 
V.
,
Belov
 
K. V.
,
Hallinan
 
G.
,
Kocz
 
J.
,
Kulkarni
 
S. R.
,
McKenna
 
D. L.
,
2020
,
Nature
,
587
,
59
 

Bolton
 
J. G.
,
Slee
 
O. B.
,
1953
,
Aust. J. Phys.
,
6
,
420
 

Bondonneau
 
L.
,
Grießmeier
 
J. M.
,
Theureau
 
G.
,
Bilous
 
A. V.
,
Kondratiev
 
V. I.
,
Serylak
 
M.
,
Keith
 
M. J.
,
Lyne
 
A. G.
,
2020
,
A&A
,
635
,
A76
 

Bridle
 
A. H.
,
Purton
 
C. R.
,
1968
,
AJ
,
73
,
717
 

Burns
 
J. O.
,
2020
,
Phil. Trans. R. Soc. A
,
379
,
20190564
 

Cane
 
H. V.
,
1979
,
MNRAS
,
189
,
465
 

Caswell
 
J. L.
,
1976
,
MNRAS
,
177
,
601
 

Chen
 
X.
,
Yan
 
J.
,
Deng
 
L.
,
Wu
 
F.
,
Wu
 
L.
,
Xu
 
Y.
,
Zhou
 
L.
,
2020
,
Phil. Trans. R. Soc. A
,
379
,
20190566
 

Colgate
 
S. A.
,
Noerdlinger
 
P. D.
,
1971
,
ApJ
,
165
,
509
 

Cong
 
Y.
,
Yue
 
B.
,
Xu
 
Y.
,
Huang
 
Q.
,
Zuo
 
S.
,
Chen
 
X.
,
2021
,
ApJ
,
914
,
128
 

Ellis
 
G. R. A.
,
1957
,
J. Geophys. Res.
,
62
,
229
 

Ellis
 
G. R. A.
,
1962
,
Nature
,
193
,
258
 

Ellis
 
G. R. A.
,
1965
,
MNRAS
,
130
,
429
 

Ellis
 
G. R. A.
,
Hamilton
 
P. A.
,
1966
,
ApJ
,
146
,
78
 

Feng
 
Y.
 et al. ,
2024
,
ApJ
,
974
,
296
 

Garland
 
C. W.
,
Nibler
 
J. W.
,
Shoemaker
 
D. P.
,
2009
,
Experiments in Physical Chemistry, 8th edn
.
McGraw-Hill
,
New York
, p.
119

Getmantsev
 
G. G.
,
Karavanov
 
V. S.
,
Korobkov
 
Yu. S.
,
Tarasov
 
A. F.
,
1969
,
Sov. Astron.
,
12
,
743

Hartz
 
T. R.
,
1964
,
Nature
,
203
,
173
 

International Reference Ionosphere
,
2022
, available at:
https: irimodel.org (accessed November 2022)

Jacchia
 
L. G.
,
1970
,
Smithsonian Astrophysical Observatory Spec. Rep
. p. 
313

Kaiser
 
M. L.
,
2005
,
Adv. Space Res.
,
36
,
1483
 

Kasper
 
J.
,
Lazio
 
J.
,
Romero-Wolf
 
A.
,
Lux
 
J.
,
Neilsen
 
T.
,
2019
,
IEEE Aerospace Conf.
,
2019
,
1

Kasper
 
J.
,
Lazio
 
T. J. W.
,
Romero-Wolf
 
A.
,
Lux
 
J.
,
Neilsen
 
T.
,
2022
,
IEEE Aerospace Conf.
,
2022
,
1

Kassim
 
N. E.
,
Weiler
 
K. W.
, eds.,
1990
,
Low Frequency Astrophysics from Space
.
Springer
,
Berlin
 

Krishnakumar
 
M. A.
,
Mitra
 
D.
,
Naidu
 
A.
,
Joshi
 
B. C.
,
Manoharan
 
P. K.
,
2015
,
ApJ
,
804
,
23
 

Lazio
 
T. J. W.
,
Cohen
 
A. S.
,
Kassim
 
N. E.
,
Perley
 
R. A.
,
Erickson
 
W. C.
,
Carilli
 
C. L.
,
Crane
 
P. C.
,
2006
,
ApJ
,
642
,
L33
 

Li
 
C.
 et al. ,
2021
,
Space Sci. Rev.
,
217
,
35
 

Lorimer
 
D. R.
,
Kramer
 
M.
,
2005
,
Handbook of Pulsar Astronomy
.
Cambridge Univ. Press
,
Cambridge

Lorimer
 
D. R.
,
Yates
 
J. A.
,
Lyne
 
A. G.
,
Gould
 
D. M.
,
1995
,
MNRAS
,
273
,
411
 

Lorimer
 
D. R.
,
Bailes
 
M.
,
McLaughlin
 
M. A.
,
Narkevic
 
D. J.
,
Crawford
 
F.
,
2007
,
Science
,
318
,
777
 

Pastor-Marazuela
 
I.
 et al. ,
2021
,
Nature
,
596
,
505
 

Phillips
 
J. A.
,
Wolszczan
 
A.
,
1989
,
ApJ
,
344
,
L69
 

Pignalberi
 
A.
,
Pezzopane
 
M.
,
Nava
 
B.
,
Coïsson
 
P.
,
2020
,
Sci. Rep.
,
10
,
17541
 

Pleunis
 
Z.
 et al. ,
2021
,
ApJ
,
911
,
L3
 

Prol
 
F. S.
,
Smirnov
 
A. G.
,
Mainul Hoque
 
M.
,
Shprits
 
Y. Y.
,
2022
,
Sci. Rep.
,
12
,
9732
 

Rajan
 
R. T.
,
Boonstra
 
A. J.
,
Bentum
 
M.
,
Klein-Wold
 
M.
,
Belien
 
F.
,
Arfts
 
M.
,
Saks
 
N.
,
van der Veen
 
A.-J.
,
2016
,
Exp. Astron.
,
41
,
271
 

Reber
 
G.
,
Ellis
 
G. R. A.
,
1956
,
J. Geophys. Res.
,
61
,
1
 

Roberts
 
C. E. Jr
,
1971
,
Celest. Mech.
,
4
,
368
 

Schreiter
 
L.
,
Stolle
 
C.
,
Rauberg
 
J.
,
Kervalishvili
 
G.
,
van den Ijssel
 
J.
,
Arnold
 
D.
,
Xiong
 
C.
,
Callegare
 
A.
,
2023
,
Radio Sci.
,
58
,
e2022RS007552
 

Shkolnik
 
E. L.
,
2018
,
Nat. Astron.
,
2
,
374
 

Spitzer
 
L.
,
1962
,
Physics of Fully Ionized Gases, 2nd edn
.
Wiley-Interscience
,
New York

Stanislavsky
 
L. A.
,
Bubnov
 
I. N.
,
Konovalenko
 
A. A.
,
Stanislavksy
 
A. A.
,
Yerin
 
S. N.
,
2023
,
A&A
,
670
,
A157
 

STEREO
,
2022
, available at:
http://en.wikipedia.org/wiki/STEREO (accessed November 2022)

Sundkvist
 
D. J.
 et al. ,
2016
,
AGU SH11C-2271
.
American Geophysical Union
,
Washington, DC

Usov
 
V. V.
,
Katz
 
J. I.
,
2000
,
A&A
,
364
,
655

WAVES
,
2020
, available at:

Weber
 
R. R.
,
Alexander
 
J. K.
,
Stone
 
R. G.
,
1971
,
Radio Sci.
,
6
,
1085
 

Wikipedia
,
2022a
, available at:

Wikipedia
,
2022b
, available at:

Yan
 
J.
 et al. ,
2023
,
Exp. Astron.
,
56
,
333
 

Zhang
 
Y.-K.
 et al. ,
2023
,
ApJ
,
955
,
142
 

APPENDIX A: TRANSIENT AND VARIABLE SOURCES

Observation of FRB at frequencies |$\lesssim 10\,$| MHz would constrain their radiation mechanisms and environments. Dense plasma prevents the escape of low frequency radiation, but at least one FRB is known to have a very clean local environment (Zhang et al. 2023; Feng et al. 2024). Old neutron stars, dead at higher frequencies, might pulse in the 1–10 MHz band. However, dispersion and scatter broadening may preclude the observation of transient sources (such as FRB) and rapidly varying sources (such as PSR) at frequencies |$\sim 3\,$| MHz. In this appendix we estimate the obstacles.

A1 Dispersion and broadening

Dispersive time delays are large at low frequencies (where the dispersion measure DM has been scaled to convenient values for FRB):

(A1)

and

(A2)

For an assumed |$\Delta \nu = 0.05\nu$|⁠, the bandwidth of a dipole antenna,

(A3)

An impulsive 3 MHz signal would be broadened to several hours after propagating through a dispersive medium characteristic of FRB or even of the more distant Galactic pulsars, and to tens of s at the plausible |$\text{DM} \sim 0.3\!-\!1 \,$| pc-cm|$^{-3}$| of the closest neutron stars.

De-dispersion is limited by the sensitivity (equation A2) of the delay to frequency, requiring |$\sim [({\rm d}\Delta t/{\rm d}\nu )B]/\delta t$| frequency channels to de-disperse a pulse of length |$\delta t$|⁠; for a hypothetical 1 ms FRB this is |${\cal O}(10^7)$|⁠. Because the bandwidth B is limited by the antenna to |$\approx 1.5 \times 10^5\,$| Hz, this would require |$\sim 100\,$| s of observation, and it would not be possible to distinguish any |$\delta t \lesssim 100\,$| s from an impulse.

In addition to dispersion, multipath scattering broadens transients (Krishnakumar et al. 2015). This is a large effect at low frequency because it scales approximately as the |$-4$| power of frequency, suggesting broadening of FRB to widths of order a day. Scatter broadening is a diffusive process that generally cannot be deconvolved; even if the scattering Green’s function were known, deconvolution would be unstable. Hence scattering likely precludes observations of most millisecond transients at 3 MHz. However, observed scattering measures differ by several orders of magnitude among Galactic pulsars, and may be very small for nearby objects, such as the closest neutron stars.

Dispersion and multipath scattering between the source and the Solar System do not affect the geometrical time delay |$\delta t$| between the direct and the ionospherically reflected signals (the ionospheric dispersion measure from the altitude at which 3 MHz radiation is reflected to deep space is only |$\sim 10^{-6}\!-\!10^{-7} \,$| pc-cm|$^{-3}$|⁠). Hence the autocorrelation of the signal received by an antenna in LEO would show a peak at the geometrical time delay |$\delta t$| (equation A4).

This peak would be broadened by the dispersive time delay (equation A1). As a result, the autocorrelation of a signal of width |$\tau$| at its source would be broadened by |${\cal O}(\tau /\Delta t)$|⁠, and the signal-to-noise ratio of detection of the autocorrelation peak would be reduced by |${\cal O}(\sqrt{\tau /\Delta t}) \sim 10^{-4}$| for the dispersion measure of a cosmological or even distant Galactic plane source (the observed Galactic FRB 200428 had |$\text{DM} = 332.7\,$| pc-cm|$^{-3}$|⁠; Bochenek et al. 2020). However, FRB 200428 had an observed L-band fluence |$1.5\,$| MJy-ms, about a million times greater than that of a typical cosmological FRB detected with |$S/N \ge 10$|⁠. Extrapolation to |$\lt 10\,$|MHz is speculative, and pulsars show spectral turnovers: detection of the autocorrelation peak is uncertain but conceivable.

As the antenna moves in its orbit, a source’s zenith angle |$\theta$| and autocorrelation peak |$\delta t$| vary. As discussed in Section A3, with knowledge of the orbit the time dependence of an autocorrelation peak may be measured, indicating the presence of a variable source as well as its location on the sky; only one orbiting antenna would be required. Additional positional information may be obtained from the times of occultation by the Earth’s limb.

A2 Detecting transients?

The autocorrelation of the baseband voltage peaks at a lag corresponding to the delay between the direct and ionospheric-reflected paths and may enable the detection of even heavily broadened and dispersed transients. This is only possible for an antenna in LEO, for which the direct and reflected signals have similar strength. A few pulsars are known with |$\text{DM} \lt 3$| pc-cm|$^{-3}$| (Phillips & Wolszczan 1989; ATNF Pulsar Catalogue 2005), implying little dispersion and scatter broadening. More numerous and even closer old and slow pulsars, ‘dead’ at higher frequencies, might have even smaller dispersion and scatter broadening. Even if de-dispersion is not feasible, pulsars may be detected from their averaged radiation (Andrew et al. 1964).

It is not known if pulsars emit radiation in the 1–10 MHz band, but pulsars with periods from 0.25 to 1.27 s have been detected at frequencies as low as 25 MHz (Phillips & Wolszczan 1989), demonstrating their emission at low frequencies and suggesting that detectable emission may extend to even lower frequencies. Detection would constrain PSR radiation mechanisms because radiation frequencies are determined by the energy of the radiating particles, the local magnetic field, and the plasma processes by which they emit; detection at lower frequencies would give different numerical values to these constraints. Detection would also provide new information about PSR population statistics and spindown history.

Coherent emission from FRB has been detected at frequencies as low as 110 MHz (Pleunis et al. 2021) and 120 MHz (Pastor-Marazuela et al. 2021), with no evidence of a low frequency turnover or cutoff. This does not demonstrate the observability of FRB at 3 MHz, but either detection or non-detection would put bounds on processes in FRB. Colgate & Noerdlinger (1971) suggested coherent supernova emission at low frequencies and Usov & Katz (2000) suggested coherent radio emission by gamma-ray bursts.

For a receiver bandwidth of 0.15 MHz, 5 per cent of the frequency, and an integration time of 3 s (a nominal pulsar with a duty cycle of 0.03 observed for 100 s); |$Bt_{\rm int} \approx 5 \times 10^5$| and |$F_{\rm thresh} \approx 5 \times 10^4\,$| Jy. This is equivalent to equation (A1.21) of Lorimer & Kramer (2005) for the mean detectable flux if two orthogonal polarizations are observed, noting that equation (3) refers to the flux density during the pulsar’s pulses while Lorimer & Kramer (2005) refer to the mean. The sensitivity may be further improved by the square root of the number of pulses in the data by coherent processing at a hypothetical or known pulsar period, but if the period is not known a priori |$S/N$| must be multiplied by 1 plus the natural logarithm of the number of trial periods. For an unbroadened 1 ms (after de-dispersion) FRB |$Bt_{\rm int} = 1.5 \times 10^2$| and the detection threshold increases to |$\approx 3.5 \times 10^6 [10 + \ln {(t_{int}/1\, \text{ms})}]/10\,$| Jy to allow for the number of possible times when the FRB might occur. This is of the same order as the observed flux of the Galactic FRB 20200428 (Bochenek et al. 2020), so future Galactic FRB may be observable.

The flux densities of coherent radio sources may increase rapidly with decreasing frequency (because they are non-thermal they need not be self-absorbed), and are expected to be much greater at 3 MHz than in L-band. Pulsars typically have 100 MHz–1 GHz spectral indices |$\sim -1.6$| (Lorimer et al. 1995), although measurements at frequencies below 100 MHz find spectral breaks (Bilous et al. 2020; Bondonneau et al. 2020). In the absence of a break, extrapolation would suggest flux densities |$\sim 10^4$| times higher at 3 MHz than at 1 GHz, or |$\sim 10^{10}\,$| Jy for a source like FRB 20200428 (a magnetar, rather than a radio pulsar, but also a non-thermal emitter). The estimated detection threshold at 3 MHz of |$\sim 3.5 \times 10^6$| Jy is much less that its (speculative) extrapolated 3 MHz flux density. However, scatter broadening, unlike dispersion, cannot be deconvolved, so these estimates may be unduly optimistic.

A3 Localization by ionospheric reflection

The ionosphere is a good reflector at 3 MHz, so the antenna would observe the reflection of a transient by the ionosphere as well as the direct signal. A flat-ionosphere approximation is justified because for the suggested orbit the antenna is a height |$h \sim 300\,$| km above the reflecting layer. That layer is approximately spherical with a radius |$R \approx 7000\,$| km, about 600 km above the Earth’s surface. If the electron density contours are smooth and horizontal, the difference |$\delta \ell$| between the direct and reflected signal paths is determined by the source’s zenith angle |$\theta$| and elevation |$\Delta = \pi /2 - \theta$|⁠:

(A4)

If the signal is a brief pulse, a dipole antenna, with broad angular acceptance, would observe two pulses separated by a time interval |$\delta t \approx \delta \ell /c$|⁠, localizing the source to a circular arc on the sky. Simultaneous detection by two telescopes would confine the source location to the two intersections of two arcs. If the source fluctuates then the autocorrelation of its base-band signal may have a peak at the lag |$\delta t$| even if the emission extends over a time |$\gt \delta t$|⁠. If natural transient sources are not detectable, an artificial orbiting beacon could be used, although beyond the scope of the minimal antenna system.

Use of equation (A4) requires knowing the instantaneous height of the reflective layer, which varies with Solar activity. It can be measured in real time if the antenna emits a pulse and receives its reflection. Rapidly varying or impulsive sources within the Solar System, anthropogenic or natural, may be localized by this method. In fact, the reflective surface of the topside ionosphere may not be accurately flat because of ionospheric turbulence, but may vary in an uncertain manner, limiting the accuracy of this method of source localization.

Transients and rapidly varying sources are advantageously observed from above the ionosphere at frequencies below the ionospheric cutoff because of shielding of terrestrial electromagnetic interference and because in LEO the delay between direct and ionospherically reflected signals constrains their direction. The peak of the autocorrelation of the baseband signal at this delay determines a source’s zenith angle even without angular resolution. Measurement from an orbiting spacecraft of the zenith angles of two separate transients from the same source would determine its location on the sky. The angular accuracy of this method is about |$c \Delta t/\Delta h$|⁠, where |$\Delta t$| is the width of a transient and |$\Delta h$| the height of the satellite above the reflecting layer of the ionosphere. For a 1 ms transient this may be several tens of degrees, likely good enough to exclude many possible candidate identifications.

This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs License (https://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial reproduction and distribution of the work, in any medium, provided the original work is not altered or transformed in any way, and that the work is properly cited. For commercial re-use, please contact [email protected]