Abstract
This paper presents a novel concept for orbit determination and time synchronization of a lunar radio navigation system. The proposed approach is based on small ground antennas that simultaneously track the entire constellation using K-band frequency links, implementing the concept of multiple spacecraft per aperture. This configuration ensures sufficient data rates and provides high accuracy in Doppler, range, and single-beam interferometry observables, enabling a precise orbit determination. We assess the achieved time transfer accuracies using both the standard asynchronous two-way satellite time and frequency transfer and a novel time transfer method that leverages onboard code epoch time-stamping and precise spacecraft range information. We propose a structure for the navigation message as well as a reference frame and associated time scale for user positioning. We complete the analysis by estimating the attainable accuracies of the signal-in-space error.
- clock synchronization
- lunar radio navigation systems
- lunar reference frames
- lunar time
- microwave tracking
- multiple spacecraft per aperture
1 INTRODUCTION
The large effort undertaken by the United States’ National Aeronautics and Space Administration (NASA), the European Space Agency (ESA), and other space agencies toward crewed and robotic lunar exploration calls for an adequate navigation and positioning infrastructure in support of the many missions foreseen in the mid-term future. ESA’s answer to this need is the Moonlight initiative (Giordano et al., 2021), which aims at the development of a reliable, yet simple, lunar communication and navigation service (LCNS). The Lunar Radio Navigation Service (LRNS), one of the two services provided by Moonlight, is a smaller-scale replica of a terrestrial global navigation satellite system (GNSS) in the lunar environment. The pillars of such a system include an accurate definition of the reference frames and time scales, precise satellite ephemerides, time synchronization across the constellation and with the ground control center, stable onboard clocks, and suitably constructed one-way radio signals to be delivered to the end user, together with a precise navigation message. The initial deployment of the Moonlight constellation entails four satellites in elliptical lunar frozen orbit (ELFO), offering good coverage of the southern polar region (Giordano et al., 2021).
This work, funded by ESA, was carried out by a consortium (named ATLAS) exploiting synergies between academic institutions and industries in Europe. We tackled the problem of orbit determination and time synchronization (ODTS) of the constellation nodes with an architecture based on the use of small ground antennas (approximately 30 cm in diameter) and a radio system operating at the K-band (22.55/23.15 GHz uplink, 25.5/27.0 GHz downlink), exploiting the concept of multiple spacecraft per aperture (MSPA) (Amin, 2016). The radio link is based on two-way spread spectrum (SS) signals capable of providing both typical range and range-rate measurements with accuracies of approximately 30 cm and 0.01 mm/s, respectively, at 10- and 60-s integration times, and single-beam interferometry (SBI) observables, a novel data type discussed in Section 3.3 (Gregnanin et al., 2012). This combination allows the generation of precise ephemerides with root mean square position errors well below the specification of the system (25 m in the initial phase of system deployment, 10 m in the fully deployed configuration), expressed in terms of signal-in-space error (SISE; see Section 5). When complemented by an accurate model of spacecraft dynamics, the aging of the ephemerides is slow, requiring an update period not shorter than a few hours. Time synchronization at the level of 1–5 ns between the satellite clocks (rubidium atomic frequency standard [RAFS] or miniRAFS) and the constellation reference clock can be attained by taking advantage of the technology currently under development by Thales Alenia Space Italy (TAS-I) for the K-band transceiver, a crucial unit for communications to/from Earth for such lunar missions. This unit is derived directly from the TAS-I integrated deep space transponder (IDST), developed for the NASA mission VERITAS to Venus (Gelfusa et al., 2019) and already equipped with the necessary technology, particularly the digital platform capable of managing the SS signal with a chip rate on the order of 25 Mcps. Time transfer is enabled by exchanging code epoch time stamps both in the standard asynchronous mode, as in two-way satellite time and frequency transfer (TWSTFT; Howe et al., 1989), or in synchronous mode with the aid of range information provided by the orbit determination (OD) process. In both modes, uplink commands and downlink telemetry do not interfere with the time transfer function, but the synchronous mode has the crucial advantage of full compatibility between simultaneous two-way radiometric (ranging and Doppler) measurements and clock comparison operations. Hence, the synchronous mode is recommended over the widely adopted asynchronous time transfer mode.
In summary, the proposed architecture rests on several novel technical solutions that provide considerable advantages over traditional lunar mission control and tracking. The pillar of this concept is the adoption of MSPA, which allows nearly continuous contact between each satellite of the constellation and at least one ground station, enabling, in addition to ODTS functions, significant data volume exchange across the radio link (telecommands, housekeeping, navigation message, etc.). The MSPA approach is associated with SS modulation and code-division multiplexing with majority voting (CDM-M) (the latter has already been tested at the engineering model (EM) level). The use of small antennas, needed to keep the entire constellation in the field of view of the ground antenna, simplifies the deployment of the ground infrastructure, while granting sufficient data volume for housekeeping and control of the constellation. The proposed architecture allows nearly complete separation of the satellite OD from the time synchronization within the constellation. This separation is made possible by the use of two-way radio links, providing radiometric observable quantities (range, range rate, and SBI) that do not rely on onboard clocks. The synchronous time transfer method described in Section 4 enables nearly continuous clock comparisons without interfering with the acquisition of the radiometric observables, with a significant improvement in the characterization and monitoring of the onboard clocks. Finally, the current paper presents results for a K-band trunk link. Although the K-band has many advantages over the X-band (as discussed in Section 2), all technical solutions are also possible in the X-band, although with significant limitations in terms of accuracy and expandability.
Finally, we wish to highlight the difference between the architecture proposed in this paper and previous work available in the literature. Lunar radio navigation and communication systems have been addressed in different ways by several authors. Some concepts exploit a combination of relay satellites and beacons on the lunar surface (Schier, 2007), without considering a full GNSS-like constellation. In some proposed implementations, the trunk link plays a minor role, as the OD function is primarily carried out autonomously by the satellites (Israel et al., 2020). The details of this innovative architecture are not yet known to the level needed for a complete comparison with the design proposed in this paper for the Moonlight LCNS. Other recently published concepts rely on traditional tracking of the lunar constellation from ground antennas (Stallo et al., 2022), without the use of MSPA and CDM-M.
2 RADIO SYSTEM ARCHITECTURE
The radio system architecture for telemetry, tracking, and command (TT&C) of the constellation during nominal operations, as well as for ODTS, hinges on two main design choices, driving LCNS performance, signal structure definition, and ground and onboard hardware.
The first choice is the use of the K-band for uplink and downlink signals, offering numerous advantages compared with the X-band. The most obvious advantage is the larger available bandwidth (500 MHz overall, compared with 50 MHz for the X-band), allowing for a higher code chip rate in SS modulation and increased accuracies of range measurements and time transfer. The K-band is also much less affected by ionospheric delays. Yet, the strongest advantage of the K-band lies in the expandability of the constellation, which is problematic in the congested X-band.
The second design choice concerns the use of a CDM scheme based on SS modulation, which enables the same frequency band to be allocated to the four satellites of the constellation with a consequent saving of spectral resources, at the expense of additional design complexity for the ground station modem and the onboard transponder. Furthermore, if the CDM technique is implemented in an MSPA scheme, it is possible to track, control, and synchronize the entire constellation simultaneously by implementing nearly continuous TT&C operations and parallelizing OD and time transfer functions (Figure 1). Because MSPA requires the constellation to be within the main lobe of the ground antenna, the constellation geometry drives the maximum size of the ground terminal. A pointing strategy based on the minimum enclosing circle curtails most pointing losses. Analyses show that a half-power beamwidth (Iess et al., 2023) of approximately 3° guarantees 99% visibility for the full constellation, corresponding to a dish diameter of 26 cm at the K-band. Although small antennas are inexpensive and easily maintained, they are also more exposed to radiation disturbances from the surrounding environment, necessitating a resilient signal structure, as presented in the next section. However, it is well known that SS receivers have intrinsic interference rejection properties, with longer SS sequences exhibiting better interference suppression.
MSPA architecture, showcasing the capability to simultaneously perform TT&C operations; Doppler, range, and SBI measurement acquisition; and time transfer for all four spacecraft of the constellation
2.1 Signal Structure
SS modulation represents the backbone of any GNSS, as it enables simultaneous signal transmission, better time-delay resolution, and good immunity to interference, especially for long SS sequences (Meurer & Antreich, 2017). The ground station transmits 24-Mcps SS signals with a modulation format derived from work by the Consultative Committee for Space Data Systems (CCSDS, 2011, 2013). The modulation is offset quadrature phase-shift keying (OQPSK) with a CDM scheme, where the four SS streams associated with the in-phase (I)-channel are superimposed and then fed to the majority vote logic (CDM-M) used to generate a constant-envelope binary phase-shift keying (BPSK) signal, suitable for operations in nonlinear channels (Donà & Iess, 2017; Donà & Rovelli, 2016; Campa, 2022). For a general overview on SS and code-division multiple access (CDMA), see, for example, the work by Holmes (1982).
The four codes of the I-channel are short Gold codes (210 – 1 chips in length) that are also used for onboard carrier signal acquisition. The quadrature (Q)-channel code (same for all four satellites) is a truncated maximum length (TML) sequence generated by a linear feedback shift register with N = 18. The TML sequence has a length of 256 times the length of the Gold code [256 (210 – 1) chips] and is synchronous with the Gold code. This feature is used by the onboard transponder to aid in acquisition of the long TML sequence. After the onboard signal acquisition, the ground station can send telecommand data on the relevant I-channel component.
The return link signal generated by the onboard transponder is different depending on the transponder configuration (Table 1), i.e., whether it is in coherent or non-coherent mode. In non-coherent mode, short Gold codes are transmitted to simplify the implementation of the SS acquisition algorithm at the ground station and to acquire the relevant quadrature phase-shift keying (QPSK) telemetry. In coherent mode, two TML sequences (of the same family as the received sequence) are transmitted in the I- and Q-channels. The two downlink TML sequences are synchronized in terms of both chip rates and code epoch with the uplink TML sequence. In this way, the ground station modem can acquire the long TML sequences aiding the SS code search with a preliminary knowledge of the spacecraft distance and the associated turn-around delay. The TML sequence turn-around in coherent mode allows the ranging, Doppler, and code epoch measurements to occur, parallelizing ODTS functions. Finally, QPSK modulation enables a balanced I and Q modulation format (telemetry data on both I and Q components).
Forward link modulation format based on the CDM-M scheme
Note: TC: telecommand; UQPSK: unbalanced quadrature phase-shift keying
This signal structure is not considered adequate for asynchronous time transfer (i.e., with the transponder in non-coherent mode). For this scenario, the proposed solution (not addressed by CCSDS (2011, 2013)) is based on the downlink transmission of the same type of modulation format and codes specified for the uplink. In detail, the proposed solution includes a short Gold code (length of 210 – 1 chips) on the I-branch and a TML sequence on the Q-branch that has a length of 256 times the length of the short Gold code [256 (210 – 1) chips] and is synchronous with the Gold code. This feature is used by the ground station to aid in acquisition of the long TML sequence. For the uplink, the modulation is unbalanced OQPSK (I/Q power ratio of 10 dB), with telemetry modulation on the I-branch and Q-branch and carrying the TML sequence, used for time transfer operations only.
In some mission phases, for instance, during contingencies, the TT&C link can be configured with pulse code modulation (PCM)/phase-shift keying (PSK)/phase modulation (PM) CCSDS standards (CCSDS, 2021), which significantly widens the ground support beyond the stations dedicated to the LRNS. The satellite in contingency is contacted by a dedicated ground station equipped with a high-directivity (and power) antenna to close the link via an onboard low-gain antenna (LGA). Although the interference between the standard-modulated signal (for the contingent satellite) and the SS signals (for the nominal satellites) must be kept under control, the power unbalance favors the former, which, coupled with the limited interference suppression capability of short Gold codes, suggests the use of mitigation strategies. One option is to employ a different frequency channel for the contingency mode. This approach would require a frequency-flexible transponder, a feature that could also facilitate simultaneous launches by leveraging external ground stations supporting only CCSDS standard modulation during the launch and early orbit phase (LEOP). Alternatively, a proper filtering mechanism in the SS acquisition/demodulation processing could be implemented to suppress the standard signal interference, as proposed by Simone et al. (2010), who analyzed an adaptive cancellation algorithm for jamming rejection in SS receivers.
2.2 Onboard Radio System
Figure 3 shows the proposed architecture for the onboard radio system, which consists of the following:
Onboard radio system architecture
The configuration during nominal operations entails the main branch (Transponder_1, Amplifier_1, and Diplexer_1) connected to the MGA, with the redundant chain (Transponder_2, Amplifier_2, and Diplexer_2) connected either to LGA_1 or LGA_2. Waveguides are indicated as black lines, whereas the red lines indicate connections that can be implemented as coaxial cables to conserve mass. RX: receiver; TX: transmitter
• One 30-cm steerable, two-degree-of-freedom, medium-gain antenna (MGA) for nominal operations
• Two LGAs for quasi-omnidirectional coverage when the MGA is not available (as in LEOP or contingency phases)
• Two redundant K-band dual-mode transponders (hot redundancy for the receiver and cold redundancy for the transmitter)
• Two traveling-wave-tube amplifiers (TWTAs; 10 W) in cold redundancy
• The radio frequency distribution assembly (RFDA), including two diplexers and five switches for signal routing to/from antennas
• Waveguides (in black) to minimize radio frequency (RF) path losses, while coaxial cables (in red) can be used on the transmitter side before signal amplification
This solution is robust, as the mission is not compromised in the case of failure of one transponder or TWTA.
The presence of the coupler in the transmitter chain prior to the 10-W amplification allows easy cross-strapping between the transponders and amplifiers, improving the system reliability. Furthermore, the RFDA allows for the following:
• During nominal operations, the RFDA ensures that the nominal transponder is connected to the MGA, whereas the redundant transponder (receiver-ON and transmitter-OFF) is connected to one of the two LGAs and is ready for telecommand demodulation.
• During contingency scenarios, the RFDA keeps each transponder connected to one of the two LGAs in order to ensure near-omnidirectional coverage for ground station contact.
• The RFDA minimizes the RF losses and the impacts on the link budget through the use of switches instead of couplers.
The K-band TT&C transponder is a flexible unit capable of working in CCSDS PCM/PSK/PM standard modulation and in SS mode, interfacing with the following:
• The onboard computer (OBC) for sending demodulated telecommands and receiving the telemetry data stream to be modulated and downlinked to the ground station
• The signal generation unit (SGU), for pseudonoise (PN) code epoch time-stamping and frequency synchronization with the external ultra-stable frequency reference (EUFR), as required by the time transfer implementation
Transponders also feature their own internal oscillator, typically an oven-controlled crystal oscillator (OCXO), for TT&C operations when the EUFR is not required/available. Nonetheless, synchronization of the transponder frequency with the EUFR is highly recommended in SS mode to reduce the signal acquisition frequency range that would otherwise suffer from the OCXO frequency instability; in addition, this synchronization is considered indispensable for the asynchronous time transfer function.
2.3 Technology Survey and Technology Readiness Level
The proposed radio system architecture makes use of several enabling features:
• The K-band instead of the more commonly used S- or X-band
• The use of wideband signals, with the current baseline considering an SS chip rate in the range of 24 Mcps
• SS modulation based on the CDM-M scheme
• Ground station modulators and receivers enabling MSPA
• Dual-mode transponders working with CCSDS standards and SS modulation
• Interference suppression onboard to handle the CCSDS standard signal in the contingency scenario and, if needed, at the ground station to avoid link interruption due to the limited rejection of SS codes
• Transponder frequency flexibility, recommended for LEOP in case of multiple launches
TT&C technology is well proven and has been demonstrated in-flight, even at the higher Ka-band frequency, as shown by the BepiColombo and JUICE missions (Cappuccio, Notaro et al., 2020; Iess et al., 2021). Their transponders, built by TAS-I, enable the novel 24-Mcps PN modulation scheme, providing two-way range measurements with centimeter-level accuracy at an integration time of only 4 s at 0.3 AU. In tracking passes with favorable weather conditions, the range-rate measurements attained an average accuracy of 0.01 mm/s for a 60-s integration time. Dual-mode transponders (compatible with CCSDS standards and SS modulation) implementing flight frequency flexibility have already been developed for the TT&C link in terrestrial GNSS constellations. The CDM-M scheme has been proposed, studied, and implemented at the breadboard level for onboard application under ESA contract, for both spacecraft (Donà & Iess, 2017) and ground station equipment (ESA-GRST-STU-SOW-0010, 2016).
Algorithms for interference cancellation in SS receivers, based on either adaptive filtering (Simone et al., 2010) or fast Fourier transform solutions, have been proposed and successfully tested at the breadboard level and have now been implemented in TT&C secure transponders. The MSPA, once supported by CDM-M technology, is primarily employed as a new approach for the control and operation of all satellites in a constellation, implementing simultaneous TT&C operation (with telecommand, telemetry, and ranging) for all links with visibility from the ground station.
In the framework of the European System Providing Refueling, Infrastructure and Telecommunications (ESPRIT) project for the upcoming lunar gateway, TAS-I is developing a K-band transceiver (KBT) for communications and tracking, a nodal element of the HALO lunar communication system. The KBT is based on the IDST, a product platform under design and development by TAS-I under ESA contract. The KBT implements frequency flexibility in the range of 60 MHz in the K-band and handles wideband signals with a data rate on the order of 50 Mbps. The digital section will be fully adequate to support the 24-Mcps chip rate of the ATLAS configuration for Moonlight.
Finally, two-degree-of-freedom steerable MGAs are a consolidated space technology, used on the BepiColombo (X-band) and JUICE (dual X- and Ka-band) missions. An off-the-shelf MGA in close alignment with ATLAS ambitions is produced by the Norwegian company Kongsberg (dish diameter: 30 cm, total mass: 12 kg, power consumption: 25 W).
2.4 Link Budget and Data Rate
The primary design feature of the proposed architecture is the MSPA concept, which concurrently enables the implementation of SBI (Section 3.3). The relatively small Earth–Moon distance and the high directivity at the K-band limit the antenna size to approximately 26 cm. However, this size is sufficient to ensure adequate data rates in the radio links. Via a pointing strategy based on the minimum enclosing circle, the pointing loss from ground can be reduced to less than 3 dB for 99% of the time (1.3 dB on average).
Table 2 presents the link budget, which was computed under the assumption of conservative conditions: the link elevation is fixed at 15°, the transmission power is limited to 200 W on ground and 10 W onboard, the slant range is set to the maximum Earth–Moon distance, and the pointing losses are set to worst-case values. In addition, a room-temperature low-noise amplifier is considered on ground. Atmospheric attenuation, quoted from the DSMS Telecommunications Link Design Handbook (2015), refers to a deep space antenna site with a cumulative distribution CD = 0.9 (implying that the attenuation is ≤3 dB for 90% of the time). Choosing a dry site is important not only for attenuation (not negligible for the K-band), but also to reduce tropospheric effects degrading the quality of radiometric observables and to guarantee an electromagnetically clean site. The latter is particularly critical for small dishes, which are prone to radiation disturbances from the surrounding environment. The transmission power from ground can easily be increased by a factor of 2 or 3 (and possibly more) if enhanced uplink performance is desired. However, the onboard RF power is expected to be a limited resource, as we assumed a small-sat platform dedicated to navigation services, in line with the Moonlight concept. Owing to power and pointing requirements, a steerable MGA with a dish diameter of 30 cm has been selected for the spacecraft.
Link budget and bit rates for the ATLAS radio link configuration
The carrier signal-to-noise ratio (SNR) for the uplink and downlink is 38 dBHz and 35 dBHz, respectively, for each satellite link of the mini-constellation, including a link margin of 3 dB (suggested by the European Cooperation for Space Standardization). The uplink SNR is primarily limited by the power-sharing loss implied by CDM-M, corresponding to 8.5 dB for four (or five) spacecraft, as reported by Donà and Iess (2017). Those losses would become prohibitively large if the constellation was expanded in the future. In that case, the strategy would be to form groups of four or five satellites and to use additional ground antennas in order to simultaneously track each group. This expansion would be low-cost, given the small size of the ground terminal and the use of an existing infrastructure. The obtained signal level is considered sufficient for acquisition, provided that the onboard receiver is compatible with an acquisition window of ±130 kHz. Because of the MSPA, uplink pre-steering in the acquisition phase can only be based on a spacecraft-averaged Doppler shift owing to CDM-M. Pre-steering the carrier frequency eases the acquisition phase because each spacecraft receives a signal with a different Doppler shift (up to 150 kHz in the K-band), and each transponder has a different rest frequency, whereas the uplink carrier is the same for the entire constellation. The signal acquisition proceeds more rapidly if the uplink carrier has a frequency that minimizes, according to some cost function, the difference between the received frequency and the rest frequency for each satellite transponder. For the downlink, the ground station can adjust the frequency independently for each satellite (given the CDMA channel), making pre-steering much more effective.
Telecommand, telemetry, and ranging SNRs are computed under the assumption of the signal modulation scheme suggested by Donà and Iess (2017), namely, unbalanced QPSK with a power unbalance of 1:10 for uplink and OQPSK for downlink. A ranging precision of less than 35 cm in two-way coherent mode can be obtained using a code rate of 24 Mcps, a jitter contribution from the SNR of the uplink I-channel, and either the I- or Q-channel SNR for downlink. The uplink format contains only the long PN code used for ambiguity resolution and coherently generated from the shorter code used for acquisition on the I-channel. Conversely, for downlink, only the long code is used, both for acquisition and ambiguity resolution. Telecommand and telemetry data rates are reported for two possible coding schemes. The target bit error rate is 10−5 for telecommand and 10−6 for telemetry. Note that the MSPA concept maximizes the data volume rather than the data rate: it allows simultaneous and continuous communication with all spacecraft. For a case with roughly 22 h of tracking per day, the overall downlink data volume from each satellite is greater than 110 Mbit/day.
In contingency scenarios, the small ground antenna performance is insufficient, and a dedicated larger dish with a diameter of at least 13 m is required. In such cases, the MSPA concept is not necessary, as the link is established with a single satellite at a time, utilizing the onboard LGA and standard modulation, with a residual carrier, as recommended by the CCSDS (2021). The SS signals from a nominally operating spacecraft will be much lower in power (by approximately 34 dB) than the residual carrier from a 13-m ground antenna. Thus, their interference is expected to be negligible. Conversely, the residual carrier signal may interfere with the SS signals, particularly for the uplink, where short Gold codes are employed. This problem can be overcome by using the strategies outlined above.
3 OBSERVABLE QUANTITIES FOR ORBIT DETERMINATION
A precise OD is essential for the generation of a high-quality navigation message that is valid for a few hours without exceeding the positioning requirements of the end user. The OD process relies on the processing of radiometric observables and on a good knowledge of the non-gravitational forces acting on the satellites. For a lunar constellation, the main observable quantities are range and range rate, provided by a coherent, two-way radio link, as described in Section 2. Although not strictly necessary to meet the SISE requirements for the satellites, SBI increases the accuracy and consistency of the OD. Thus, SBI is an important enhancement of our LRNS concept.
3.1 Doppler
Doppler measurements rely on the comparison between the frequency of an uplink signal generated by a highly stable frequency standard (typically a hydrogen maser) and the received signal, coherently returned by the onboard transponder. Because the relative frequency shift is equal to the time derivative of the round-trip light-time (dRTLT/dt = Δf/f) (Thornton & Border, 2003), defining range as RTLT/c, range rate and Doppler, being proportional, are often used interchangeably. Range rate, not to be confused with the difference of two range measurements divided by the time interval, is naively interpreted as the radial velocity of the spacecraft. Doppler shifts, derived from phase measurements, are a standard output of ground back-end receivers. Different from Doppler measurements, which depend on the carrier frequency, range rate is independent of the carrier frequency. State-of-the-art deep space tracking systems based on Ka-band links provide range rate accuracies in the range 0.005−0.02 mm/s at 60 s integration times, or, in terms of the Allan deviation (ADEV) (Barnes et al., 1971) of the frequency residuals after orbital fit, 1.7−6.8 × 10−14. At longer time scales (1000 s), the ADEV of frequency residuals often reaches 5 × 10−15, corresponding to 0.0015 mm/s. Water vapor radiometers (WVR) and normal weather conditions are necessary to attain these values (Asmar et al., 2005, Iess, et al., 2018; Cappuccio, Hickey et al., 2020). Doppler is the observable of choice for positioning of orbiters within the planetocentric frame, due to its extreme sensitivity to gravity gradients. For this reason, it deserves special attention in the LRNS.
The error budget of Doppler measurements for the ATLAS radio architecture has been derived from recent data collected by BepiColombo, reported by Iess et al. (2021), and from a previous analysis by Iess et al. (2014). We adjusted thermal noise to the SNR computed in Section 2.4 and neglected the contribution of interplanetary plasma. Each error source in the Doppler error budget has its own spectral properties and a different relationship with the integration time. Because the numerical simulations were carried out using Doppler data sampled at 60 s, the error budget in Table 2 has been computed for a 60-s integration time, scaling the values reported in the literature at a different integration time as needed. Note that some error sources, such as errors in the station locations and Earth orientation parameter (EOP) and mismodeling of Earth solid tides, produce larger ADEVs at longer integration times. However, their contribution, which is roughly proportional to the integration time, does not affect the end-to-end noise budget, even at 1000 s. Thus, the assumption of white noise at 60 s made in the simulations of the OD process is valid. Finally, the very small value of the ground antenna mechanical noise (8 × 10−18) is a consequence of the small size of the dish. By comparison, Asmar et al. (2005) reported a value of 4 × 10−15 at 1000 s for NASA’s Deep Space Network antenna DSS 25, a 34-m dish. The value in Table 2 was derived by scaling linearly from the value reported by Asmar et al. (2015) and accounting for the different integration times (60 s vs. 1000 s).
3.2 Range
Range measurements are affected by random noise (thermal jitter) and systematic effects (range bias). The ranging jitter due to thermal noise depends on the available SNR (see Table 2), in both the uplink and downlink. For an SS ranging system, the jitter is obtained from the following expression of the chip tracking loop (Holmes, 1982):
1
where:
2
is the equivalent signal power over noise power spectral density, Tc is the chip length , BP is the pre-detection bandwidth, and BL is the loop bandwidth. BP is conservatively taken as 8 times the symbol rate. For the onboard transponder (telecommand based on BPSK modulation on the I-branch at a bit rate of 1.8 kbps with a code rate of r = 1/2), we assumed BP = 28.8 kHz; for the on-ground receiver (telemetry based on OQPSK modulation at a bit rate of 1.4 kbps with a code rate of r = 1/2), BP = 11.2 kHz. In downlink, the pre-detection bandwidth is derived from the telemetry symbol rate in one branch, which is assumed to be half of the total symbol rate derived from the link budget.
When the onboard loop bandwidth is much narrower than the on-ground loop bandwidth, the uplink and downlink jitters are summed as the square root of their squared values. Under this unlikely condition, the overall end-to-end one-way ranging jitter can be computed as follows (CCSDS, 2014, Section 2.7.3):
3
where c is the speed of light. In a more realistic case, when the onboard loop bandwidth is much larger than the on-ground loop bandwidth, the overall end-to-end one-way ranging jitter is independent of the onboard loop bandwidth and can be written as follows (CCSDS, 2014, Section 2.7.3):
4
In our case, we have assumed a loop bandwidth of BL_U = 5 Hz for the onboard transponder and a value of BL_D = 0.1 Hz for the ground receiver (considering range observables every 10 s).
In contrast to range-rate measurements, range observables are generally affected by a significant bias, generated by both ground and onboard electronics and antennas. Both biases can be calibrated with high accuracy by means of dedicated hardware. The transponder group delay can be monitored with accuracies of <0.1 ns by means of a calibration circuit embedded in the unit. This is the case of the BepiColombo KaT, a Ka-band transponder that is part of the MORE scientific investigation (Cappuccio, Hickey et al., 2020). The two-way group delay of the ground electronics can also be calibrated with similar accuracies, as demonstrated by Cappuccio, Hickey et al. (2020) for ESA’s deep space station DSA-3 (Malargüe, Argentina). Although the transponder remains a critical element, the delays due to all intervening elements in the end-to-end chain must be calibrated (waveguides, cables, filters, amplifiers, etc.), including possible variations with temperature.
The end-to-end range error budget is reported in Table 3. Although the use of an advanced WVR would improve the tropospheric calibration of the zenith wet delay to 0.5 cm (Linfield et al., 1996), we have assumed a more conservative value of 2 cm, in line with the accuracies provided by the radiometer installed at DSA-3 (Lasagni et al., 2023) and an elevation angle of 30°. We have also used relatively large values for the station location (3 cm), as very long baseline interferometry (VLBI) observations cannot be used with small dishes. This value is compatible with positioning via laser metrology from a fiducial point, assuming that the ATLAS antenna is located close to a large deep space tracking antenna included in the VLBI network or a GNSS station.
Doppler and Range Error Budget for the ATLAS Architecture, Expressed as the ADEV for a 60-s Integration Time and Distance Error Sampled at 10 s, Respectively
3.3 Single-Beam Interferometry
The application of radio interferometry for OD is usually realized by delta-differential one-way ranging (DDOR), where two widely separated ground antennas sequentially track a spacecraft and a quasar. In a different interferometric scheme (Bender, 1994), the phases of a two-way signal received from two spacecraft, whose directions fall within the beamwidth of a single ground antenna, are simultaneously recorded and differenced. This configuration, known as SBI, was originally considered for a network of landers to study the interior structure of the Moon (Bender, 1994; Gregnanin et al., 2012) and Mars (Gregnanin et al. 2014). The SBI observable retains information on the differential range between the two spacecraft, relative to the ground antenna (Figure 1). Common noise sources, such as instabilities in the frequency standard and antenna deformations, are completely suppressed. If the angular separation between the two spacecraft is small (approximately 2°), typical noise rejection is also effective for propagation delays due to media (troposphere, ionosphere, and interplanetary plasma). Although SBI has never been implemented or demonstrated (because of the need to uplink two coherent carriers or CDMA signals from the same deep space antenna), the generation of SBI observables is expected to be straightforward in an MSPA tracking configuration such as the one considered for ATLAS.
Gregnanin et al. (2012) analyzed the performance of a Ka-band system capable of measuring the differential phase between two landers at a fraction of a wavelength (≤0.1 mm). The considered configuration entails a network of three lunar landers separated by a baseline of approximately 1000 km. As the main errors in SBI are, to first order, proportional to the angular separation between the spacecraft, the achievable accuracy for the ATLAS constellation will be lower (approximately 1.3 mm), primarily limited by the incomplete cancellation of the tropospheric and ionospheric path delays. The SBI error budget is evaluated with the same assumptions used for the link budget (see Section 2.4). The reported values for media errors provide an order-of-magnitude evaluation based on the same approach used for the DDOR error budget in CCSDS 500.1-G2 (2019), where average values for GNSS tropospheric and ionospheric calibrations are assumed. Should a WVR be available at the ground site, the residual tropospheric path delay can be significantly reduced. Note that media errors also depend on the tracking elevation (an average pass elevation of 30° is assumed in Table 4).
SBI Error Budget
4 TIME TRANSFER AND CLOCK SYNCHRONIZATION
As for any terrestrial GNSS, synchronization among the clocks of the constellation is crucial to meet positioning requirements of the Moonlight LRNS, usually provided in terms of SISE (see Section 5.4). In the proposed architecture, the ground stations and spacecraft perform ground-to-space time transfer exploiting the two-way radio link. An important difference between terrestrial GNSS and the ATLAS LRNS is the nearly complete independence between the OD process and the clock offset estimation, attained via the use of two-way observables that do not involve the onboard clocks by any means. The time synchronization accuracy across the lunar constellation relies on three main factors:
• Ground-to-space time transfer accuracy
• Accuracy and stability of the satellite clocks
• Desynchronization between different ground tracking stations
The potentially detrimental effect of desynchronization between different ground stations is largely mitigated by the MSPA approach. Indeed, a single ground station (or multiple stations of the same complex employing the same frequency standard) can simultaneously perform ground-to-space time transfer with the whole constellation, keeping the time of the constellation internally consistent. In addition, the clocks at different ground stations can be synchronized by exchanging clock offset information immediately before and after a station handover. As we will show in the following section, accuracies in the range of 1–5 ns can be attained in the ground-to-space time transfer. If clock comparisons are carried out immediately before and after the handover by the two intervening ground stations and the constellation spacecraft, the clocks of the ground network can also be internally synchronized to the same level of accuracy (1–5 ns). Binding the network time to International Atomic Time (TAI)/Coordinated Universal Time (UTC) can be achieved via GNSS clock synchronization. Note however that user positioning in the lunar environment requires only a good synchronization within the constellation, which is not necessarily a synchronization to TAI/UTC. The time of the constellation must be properly related to the lunar reference frame, defined by laser ranging and lunar gravity. Here, we discuss two methods for ground-to-space time transfer and identify suitable clocks compatible with small spacecraft (mass ≈ 150–200 kg) and the required SISE.
4.1 Clock Synchronization Methods
Any measure of time desynchronization between distant clocks requires a comparison of their readings at the same coordinate time. This process involves relativistic transformations between proper time τ (the actual time read by each clock) and coordinate time t. However, we report only a simplified formulation here, where de facto proper and coordinate times are assumed to be equivalent. Indeed, relativistic transformations come into play only as deterministic corrections that need to be applied in an operational scenario. With this simplification, we need only to consider the finite speed of light. For a review of time and frequency transfer methods, see the work by Levine (2008).
We have identified two methods for performing ground-to-space time transfer. The first method is similar to the well-established TWSTFT asynchronous technique (Howe et al., 1989), which is routinely used on Earth for comparing distant clocks and generating the UTC time scale. The second method is based on a novel approach in which the comparison is aided by ranging measurements and OD. This synchronous technique does not require an interruption of the two-way coherent link used for navigation and TT&C, therefore allowing, in principle, a greatly increased frequency of desynchronization measurements. In addition, this second implementation is fully compliant with the SS signal, as defined by the CCSDS (2011, 2013), and, for this reason, could be adopted as the baseline solution. Ultimately, both methods rely on the combination of time tags applied onboard and on ground. The time-tagging is triggered by a specific portion of the intervening signals, called the “code epoch,” which is most easily identified as the first chip of the PN ranging sequence. The use of code epoch tagged at time of ground transmit, time of onboard receipt and transmit, and ground receipt has been employed for decades in NASA's TDRSS, although the synchronization accuracy is only at about 1 ms (Sank, 1993).
4.1.1 Asynchronous Mode (TWSTFT)
Asynchronous ground-to-space time transfer relies on non-coherent links (hence, the incompatibility with two-way radiometric measurements), similar to terrestrial TWSTFT, where geostationary satellites are used as relays for signals coming from/to remote terminals that are not in mutual visibility (ITU-R TF.1153-4, 2015). For the Moon, the two remote terminals are in mutual visibility, and the configuration is even simpler. Figure 4 shows the world lines and the scheme of the asynchronous links, along with all of the relevant epochs involved in the measurement. A ground clock at t1 generates a signal transmitted from the antenna at t2 after a certain delay in the transmitter channel, primarily due to cables and electronics. The signal reaches the satellite antenna at t3 and, after another delay in the satellite receiver channel, reaches the satellite time comparator at t4, where the onboard observable ΔτB = t5 − t4 is produced. In this scheme, t5 is the epoch at which the onboard SGU, clocked by the EUFR, generates a signal for downlink transmission, as in the uplink process. This signal is transmitted from the onboard antenna at t6 after a delay in the spacecraft transmitter channel and reaches the ground antenna at t7; after another delay in the ground receiver channel, the signal reaches the station time comparator at t8. Finally, the observable quantity ΔτA = t1 − t8 is generated at the Earth tracking station.
World lines and scheme for the asynchronous time transfer method
The two observable quantities, measured onboard and on ground, are ΔτB = t5 − t4 and ΔτA = t1 − t8, respectively. The other terms represent onboard and ground delays in the transmitter and receiver path and propagation delays (T23, T67).
Following the formulation given by Duchayne et al. (2009) and Delva et al. (2012) and combining the observables from the two asynchronous links, the desynchronization equation between the two clocks may be written as follows:
5
which shows that the clock desynchronization between ground and space clocks is the average of the two measured observables corrected for differential delays due to electronic and propagation delays (including geometric path delays and additional effects due to the Earth’s troposphere and ionosphere). If t1 ≈ t5, the path reciprocity ensures that most errors cancel out in the differentiation procedure, including errors in OD. According to the error budget, accounting for all time error contributors, the end-to-end measurement uncertainty could be as low as ∼ 0.3 ns, remarkably in line with the uncertainty reported in the Bureau International des Poids et Mesures (BIPM) Circular T (BIPM, 2023).
A drawback of the asynchronous method is the need to switch the operational mode of the onboard transponder from coherent to non-coherent, which interrupts the acquisition of radiometric data at the ground station. Therefore, the number of clock comparisons performed during a tracking pass is limited. Although a practical realization of the asynchronous method requires some changes in the hardware of state-of-the-art deep space transponders, the current digital architectures make such modifications easy to implement.
Figure 5 shows a block diagram of the onboard architecture and interfaces between the transponder, SGU, EUFR, and onboard computer (OBC) that allow implementation of the asynchronous TWSTFT method. In this mode, the transponder should be configured to operate in non-coherent mode synchronized in frequency with the EUFR.
Block diagram for the onboard architecture supporting time transfer asynchronous mode based on non-coherent links
The EUFR provides the time reference to the transponder for receiver and telecommand code epoch time-tagging. The OBC uses its internal clock for telemetry frame time-tagging, as the code epoch time-tagging has already occurred. CE: code epoch; CLK: clock; RX: receiver; TM: telemetry; TX: transmitter; XO: internal oscillator
The two separate registers inside the SGU should operate independently to record readings from the SGU counter according to separate trigger signals associated with the code epoch in the received signal and the code epoch in the deep space transponder transmitter code generator, clocked by the EUFR. These readings are collected by the OBC, where they are formatted in the telemetry data stream and sent to the ground station via the TT&C link. On ground, they are used in post-processing to generate the onboard observable ΔτB to be applied in the desynchronization equation (Equation (5)). The SGU counter clock frequency should be on the order of ∼300 MHz to provide a time-stamp resolution at the level of 3.3 ns. This resolution may then be further improved by an averaging process.
As an alternative to the standard synchronization marker (SM) approach used, for example, in the ESA GAIA mission, the proposed architecture also includes a fast strobe line (receiver code epoch) to trigger the time-stamp operation clocked by a standard crystal oscillator within the OBC. In this case, the triggering signal can be either the receiver code epoch or the SM resulting from the output of a telemetry frame generator; the latched value inside the register is the OBC spacecraft elapsed time (SCET). It is worth noting that the navigation payload time to be synchronized with UTC/TAI is referred to here as the onboard time, which is distinct from the OBC time, indicated here as SCET. The latter is less accurate because of uncertainties in the telemetry chain and/or crystal oscillator instability but could be used in mission phases when the other method cannot be used, or even exploited to initiate the time transfer algorithm for onboard time synchronization.
4.1.2 Synchronous Mode
In any two-way coherent radiometric measurement (range or Doppler), the mathematical model of the light-time considers the time of three participating events (neglecting, for now, the internal spacecraft delays):
• t1: epoch of signal transmission from ground
• t2: epoch of signal reception onboard the satellite
• t3: epoch of signal reception on ground
Two-way range observables represent a measure of the round-trip light-time (RTLT = t3 − t1). In vacuum, RTLT × c (the speed of light) provides a measure of the distance traveled by the signal in the uplink and downlink paths. The ground clock directly measures t3, whereas t2 and t1 are retrieved from a backward solution of the light-time equation (Moyer, 2003), a mathematical pillar of any OD code. Instead, the onboard clock has direct access to t2, measured in terms of its proper time and converted to coordinate time. Therefore, in a two-way coherent link configuration, the ground-to-space clock desynchronization can be accessed by comparing different measurements of the epoch t2:
6
where is the onboard time obtained from the OD solution through ρ23/c, the one-way return light-time. t2(t) are simply the readings of the onboard clock that are then transferred to ground via the telemetry data stream. This difference represents the desynchronization between ground and space clocks. The onboard time-stamping operations are triggered by a code epoch signal, activated by a specific chip of the SS signal. The proposed TML sequence for the uplink SS code has a length of 256 · (210 – 1) = 261,888 chips. Assuming a chip rate of ∼ 24 Mcps, the chip length is τ ≈ 41.6 ns, and the code repetition period is TTML ∼ 10.9 ms.
Two aspects must be carefully considered when this method is used: the capability to solve for the code ambiguity and the precise association of the code epoch arrival to the time-stamping operation carried out by the clock assembly. Data encoding and transmission through the telemetry channel may indeed add a certain delay Δ2, which, depending on the adopted coding method, may cause Δ2 to exceed TTML. However, this delay is largely deterministic and could be calibrated in post-processing, while also taking into account possible thermal effects. Therefore, the important requirement is that the residual uncertainty δΔ2 ≪ TTML, a condition that can be easily met. When the signal is received on ground, a similar time-tagging operation is triggered by the received code epoch signal, and the epoch t3 provides through the OD process. In parallel, telemetry data for t2 are downloaded by the satellite. This step may add an additional delay Δ3, for which the same considerations described for Δ2 apply. Figure 6 illustrates the transfer method based on the two-way coherent link.
Scheme for time synchronization with the two-way coherent approach
The delays Δ2 and Δ3 introduced by data encoding/decoding in the telemetry transmission may exceed the code repetition period. The associated uncertainties δΔ2 and δΔ3 must be ≪ TTML. is the onboard time of the receiver code epoch obtained from the OD solution. CE: code epoch; TLM: telemetry; RX: receiver; TX: transmitter; G/S: ground-to-space
The performances of this time synchronization method depend on a) the precision of time-stamping operations (σt2 and σt3), b) the OD accuracy in the line-of-sight direction (σρ) and c) the accuracy of the onboard and on-ground delay calibration. The Lunar Orbiter Laser Altimeter of the Lunar Reconnaissance Orbiter has achieved a time-stamping precision of approximately 0.5 ns (Bauer et al., 2017), while numerical simulations (see Section 5) show an OD accuracy better than 10 cm along the line of sight, corresponding to 0.3 ns (see Section 5.4). Therefore, we may conservatively assume an overall contribution of the ground-to-space time transfer procedure at the 1-ns level to the time synchronization across the entire lunar constellation. Note that the (much larger) errors in the other two directions (transverse to the line of sight) do not contribute to the time synchronization procedure.
In Equation (6), the contributions from the media and the internal delays at the station and outside the transponder (which need to be accurately calibrated) do not appear explicitly. The media and orbital errors are included in the term ρ23, whereas the internal calibrations also affect t2. Although appropriate media and internal delay calibrations can keep the error within a few nanoseconds, the media contributions may be strongly reduced even in synchronous mode if the uplink path is also exploited. Indeed, if Equation (6) is also written for the uplink as follows:
7
then one can combine and average two separate ground-to-space time comparisons, similar to what has been done for the asynchronous configuration. Considering the world lines in Figure 4, one may recognize that for the synchronous case t4 = t5, the combination of the two independent uplink/downlink paths leads to the following:
8
where the up and down arrows refer to the uplink and downlink geometric and media delays, respectively. The last four terms indicate the ground station delays in the transmitting and receiving chain and the onboard delays in the receiving and transmitting chains, respectively. In this way, the coherent method also exploits the benefits from differenced terms for propagation and media delays.
Figure 7 shows a block diagram of the onboard architecture and the interfaces between the onboard transponder, SGU, EUFR, and OBC. All of these units and interfaces enable the time synchronization method based on two-way coherent ranging observables, either exploiting only the downlink path (Equation (6)) or both the uplink and downlink paths (Equation (8)). We recall that in the synchronous mode, the transponder is configured to operate in two-way coherent mode. The code epoch of the TML sequence (receiver code epoch, also used for ranging turn-around) triggers a time-stamp operation by the SGU through a fast time strobe line, clocked by the EUFR. As soon as this trigger signal is received by the SGU, the reading from the SGU counter (corresponding to t2(t) of Equation (6)) is recorded in a register. This time-stamp information is routed to the OBC and incorporated into a data packet transmitted into the downlink telemetry data channel.
Block diagram for the onboard architecture supporting time transfer in synchronous mode, based on two-way coherent ranging observables
CLK: clock; RX: receiver; TLM: telemetry; TX: transmitter; XO: internal oscillator
Similar to the previous asynchronous time transfer method, the synchronous time transfer solution also includes, as an alternative to the standard SM approach, a fast strobe line (receiver code epoch) to trigger the time-stamp operation clocked by a standard crystal oscillator within the OBC. The latched value inside the register is the OBC SCET with the same functionality described for the asynchronous time transfer method.
4.1.3 Comparison of Time Transfer Architectures
As for the hardware, the two time transfer architectures described above are actually rather similar, with the TWSTFT being slightly more complex because it requires the following:
- a transponder that interfaces directly with the EUFR
- an additional register within the SGU
- an additional fast strobe line from the transponder transmitter module to the SGU
It is also worth noting that the two architectures are not mutually exclusive. De facto, the synchronous architecture may be seen as a simplified version of the asynchronous architecture.
In terms of absolute time synchronization accuracy, the two methods are expected to provide similar results, with the asynchronous time transfer mode expected to slightly outperform the synchronous time transfer owing to its better immunity to orbital errors. Because the asynchronous time transfer mode does not rely on the OD solution, it can be applied without the need for external information. Hence, it slightly outperforms the synchronous time transfer mode for reliability and quick availability of the time synchronization solution, which may constitute an advantage when one element needed for OD could be temporarily unavailable (or underperforming).
From an operational viewpoint, the main difference between the asynchronous and synchronous time transfer modes is that, in the proposed architectures, the latter technique cannot be performed during nominal tracking operations because the onboard transponder must operate in non-coherent mode, while all tracking observables (range, range rate, and SBI) require a coherent link.
A remarkable advantage of the synchronous mode is its full compatibility with the simultaneous collection of radiometric data. Hence, the synchronous mode allows much more frequent clock comparisons. Therefore, monitoring the onboard clock drifts gives more accurate results, with a beneficial impact on the predictions of the clock drifts as well.
4.2 Onboard Clocks
The selection of onboard clocks is a trade-off between performance (accuracy and stability); size, weight, and power (SWaP) and cost; and technology readiness level. Although the MSPA approach allows for nearly continuous contact between each satellite and the tracking stations, the system should be able to still work when contact with Earth is interrupted for a few hours. This requirement rules out the use of microchip-scale atomic clocks, whose drift would result in SISE values of >100 m in a few hours. Deep space atomic clocks (Burt et al., 2021) and Galileo passive hydrogen maser clocks (Steigenberger & Montenbruck, 2017) are the best choices in terms of absolute performance (especially for their long-term stability, <10−14 for a 1-day time scale) but at the cost of a mass no lower than approximately 10 kg. The long-term stability of Galileo rubidium clocks is approximately 10 times worse, but the error contribution after 1 day is still acceptable (∼ 1 m), and the mass (3.3 kg) is still compatible with use onboard a small/minisat. Two other attractive solutions are miniRAFSs and ultra-stable crystal oscillators (USOs), given their low mass and power consumption. Although USOs generally suffer from large drift, the AccuBeat USO, developed for ESA’s JUICE mission to Jupiter, combines an excellent short-term ADEV with good performance at longer time scales (∼10−12 at 10,000 s). Table 5 summarizes a SWaP and performance comparison among RAFS (Orolia), USO (AccuBeat), and miniRAFS (Orolia) (with references to manufacturer fact sheets). A rough estimation of the clock error contribution to the positioning error at different time scales may be estimated as Δp ≈ cτσy.
SWaP and Performance Comparison (ADEV) Among Different Space Clock Technologies The ADEV at different time scales is the maximum value for each technology. Data sheets for the three clocks are available at the following websites:
(1) https://safran-navigation-timing.com/product/rafs
(2) https://www.accubeat.com/uso
(3) https://safran-navigation-timing.com/product/minirafs-rb-atomic-frequency-standard
5 ORBIT DETERMINATION AND SATELLITE EPHEMERIDES
Our concept for a lunar constellation involves four satellites orbiting in ELFOs. These orbits are designed to minimize the need for orbit-keeping maneuvers while covering the southern polar regions of the Moon, key targets for future lunar missions (Melman et al., 2022). The orbit parameters are chosen to reduce the effects of gravitational perturbations, ensuring stable orbits (although inclination-dependent drifts affect the right ascension of the ascending node).
A detailed analysis of the ODTS performance of the proposed LRNS and the ground station visibility for the constellation has been provided by Sesta et al. (2025) (for a comprehensive introduction to OD, see the work by Schutz et al. (2004)). However, we present here the architecture performance for the three proposed clocks to characterize the time validity of the navigation message for the lunar constellation. In this scenario, we assume a K-band radio link between the satellites and the three ground stations (assumed to be located at the ESA deep space tracking sites) with the MSPA approach. The collected observables are Doppler, range, and SBI (with a typical sampling time of 60 s), whereas we adopt ionospheric media calibration with GNSS (10% uncalibrated) and tropospheric media calibration with an advanced WVR (5% uncalibrated). Given the nearly continuous tracking, in principle, OD solutions can be produced very frequently (e.g., at hourly rates), but the actual requirements on the frequency of the solution updates are determined by the aging of the navigation message (see the work by Sesta et al. (2025) for a detailed discussion).
Figure 8 represents the 99th percentile of the SISE as a function of the age of data (AOD) for the different clocks proposed in the previous section. The overall SISE is defined as the sum of position and clock errors:
Evolution of the 99th percentile of the SISE as a function of the AOD for miniRAFS (blue), RAFS (green), and USO (orange) clocks
The dashed red line represents the 25-m requirement for the SISE.
9
where x, y, and z are the true coordinates of the spacecraft at time t and are the broadcasted coordinates and time ( includes the clock corrections). The main contributions to the SISE are represented by ephemerides aging and clock drift. The AOD is defined as the elapsed time between the generation of a navigation message by the ground segment and its usage at the user level (European GNSS, 2021).
The relatively good long-term stability of miniRAFS and RAFS clocks allows a longer navigation message validity (because the clock error has a smaller contribution to the SISE) than the USO. Figure 9 shows that all of the proposed clocks are compatible with a 4-h message validity with a maximum SISE value of 25 m. However, with the rubidium clocks, the message update frequency can be reduced by half, guaranteeing a time window validity for the broadcasted navigation message of almost 8 h in 99% of the analyzed cases.
Maximum values and 95th percentile of the 3D orbit fit residuals for different window lengths (horizontal axis) and a varying number of Chebyshev coefficients (vertical axis) for a lunar orbit representation
All values are expressed in meters.
6 BROADCAST EPHEMERIDES REPRESENTATION
The broadcast ephemerides are typically expressed in three different ways: by using Keplerian elements with corrections, by applying Cartesian position and velocity series, or by using polynomials, e.g., trigonometric, Chebyshev, or Lagrange polynomials.
The Cartesian position series, which is applied in the GLONASS system, demands a numerical integration of the orbit based on, e.g., Runge–Kutta integration algorithms. To improve the quality of the derived positions, a full set of models accounting for orbit perturbations should be used, which limits the possible applications for real-time solutions unless a simplified model of orbit perturbations is used, owing to computational complexity on the user side. The GLONASS navigation system employs a Cartesian representation for broadcast orbits, which results in a substantial computational complexity, thus imposing a challenge for real-time solutions when compared with Galileo, BeiDou, and the Global Positioning System (GPS), which employ a Keplerian representation.
The main advantage of using polynomials, such as Chebyshev or Lagrange, lies in a simple implementation and representation of the orbit. The accuracy depends directly on the number of coefficients, i.e., the degree of the polynomial expansion. The disadvantage is the rapid growth of errors near the interval endpoints within which the polynomial is fitted. The possibilities for extrapolation outside the selected window are also limited, depending on the function and the selected data set.
Keplerian elements with corrections accounting for orbit perturbations are used in most Earth GNSSs. Corrections typically include secular rates of the parameters or once-per-revolution corrections. A standard representation based on Keplerian parameters was designed for circular orbits of high-altitude Earth satellites, for which the gravity field perturbations of Keplerian parameters can be simplified to the main terms, especially the Earth flattening J2 term. However, the orbit of a lunar satellite is notably different from that of an Earth GNSS satellite. As opposed to the nearly circular orbits of Earth GNSSs, the considered orbit of the lunar orbiter is highly eccentric. High eccentricity is needed to place the aposelene above the lunar south pole, the region of highest interest owing to the occurrence of water ice. The periselene region might be difficult to model because of its relatively low latitude above the lunar surface, which induces large gravitational perturbations caused by the lunar gravity potential, and the high speed of the satellite when compared with the aposelene region. The lunar flattening is much smaller than the Earth flattening; thus, the components accounting for the secular rate of the ascending node in Earth GNSS broadcast messages are not dominant for lunar orbiters (although they need to be considered).
Two solutions were finally considered for the lunar orbiter broadcast ephemeris representation in high-eccentricity orbits; these solutions are based on Chebyshev polynomials and Keplerian parameters with corrections. The representation using Cartesian positions and velocities was rejected because of the high computational complexity required for numerical orbit integration and the large number of bits when compared with other orbit representations.
6.1 Chebyshev Polynomials
Chebyshev polynomials constitute a sequence of orthogonal polynomials (with respect to the weight function on the interval [−1, 1]), defined as solutions of the Chebyshev differential equation (Rivlin, 2020). Chebyshev polynomials are an advantageous solution to the classic numerical interpolation problem. Their purpose in numerical analysis is to obtain well-behaved and efficiently computed approximations to smooth but complex mathematical functions. Therefore, these polynomials can be applied to represent ephemeris parameters or rotation parameters between reference frames. The recursive calculation of the Chebyshev polynomial coefficients is expressed as follows:
10
where the x values in the time domain range from −1 to 1. A final function with a larger degree of polynomial expansion and more coefficients can represent curves with higher complexity. The functions are broken up into small intervals, within which the coefficients and the resulting functions are valid. We conduct a least-squares fit of the Chebyshev polynomials into the simulated position of the lunar orbiter decomposed into X, Y, and Z coordinates. Each component of the lunar orbiter position, i.e., X, Y, and Z, is calculated as a sum of Chebyshev terms:
11
where ai denotes the coefficients from the navigation message.
Figure 9 illustrates the 95th percentile and maximum of three-dimensional (3D) residuals within a full data set covering 10 days, considering 7–12 Chebyshev coefficients (vertical axis) and window lengths of 0.5–3.0 h (horizontal axis) for the orbit of a lunar spacecraft. Chebyshev polynomials of degree six (with seven coefficients) are sufficient to obtain an orbit quality better than 1 m in a 0.5-h window. Increasing the number of Chebyshev coefficients to 10 allows for an accurate orbit representation in a window of 1 h, with the 95th percentile at the level of 0.116 m. A Chebyshev polynomial of degree 11 (comprising 12 coefficients) is sufficient to ensure an orbit reconstruction quality of up to 1 m, valid within 1.5-h intervals. In an analysis of the orbit fit time series, the primary issues arise as the satellite passes over the periselene, leading to a notable deterioration in the quality of orbit reconstruction. Employing 12 Chebyshev coefficients is sufficient to obtain a centimeter-level orbit quality in 75% of the considered 2.5-h windows, except for the windows comprising the periselene passes, where the maximum position errors reach hundreds of meters for single epochs. Assuming an orbit representation using Chebyshev polynomials with 10 coefficients and the validity of the orbit model for 1 h, the position errors do not exceed 60 cm for single epochs during the periselene passes. Therefore, this situation requires a representation with a shorter window for periselene passes than for other parts of the orbital arc. However, at the lunar south pole, signals from a satellite close to the periselene cannot be received; therefore, the periselene representation is of minor importance to fulfill the accuracy requirements. The disadvantage of the Chebyshev polynomial representation is the number of bits (187) that are needed for a single coordinate component. In total, for three position components together with time-stamp information, 595 bits are needed for an 11-coefficient representation.
6.2 Keplerian Representation
To provide a centimeter-level accuracy of the orbit recovery, a model of six Keplerian and nine empirical accelerations is recommended, including constant accelerations in the along-track, cross-track, and radial directions, as well as sine and cosine once-per-revolution corrections in this orthogonal frame. For a 1-h window, 95% of residuals are below 2.9, 5.0, and 2.9 mm in the radial, along-track, and cross-track directions, respectively. As the window increases, the quality of the orbit fit decreases. The median values of the orbit fit residual are at the level of 1, 2, 6, and 203 mm for 1-, 1.5-, 2-, and 4-h intervals, respectively. If the model is reduced to ten parameters, i.e., six Keplerian and four selected empirical parameters that provide the smallest fit error, the quality of the orbit fit is substantially reduced, i.e., the median values of the orbit fit residual are at the level of 22.8, 52.4, 91.1, and 395.0 m for 1-, 1.5-, 2-, and 4-h intervals, respectively. The Keplerian representation of six Keplerian and nine empirical parameters requires 370 bits, and another 31 bits are needed to mark the time stamp of the message, resulting in a total of 401 bits.
Determining a reasonable number of bits for the representation of the navigation message for eccentric orbits is challenging, as it requires a balance between an accurate representation of the lunar orbiter position and the storage space for communication purposes. The time required to update the navigation depends on both the constellation characteristics and the chosen orbit representation option. For Earth GNSSs, the refresh cycles for navigation messages vary from several minutes to daily. In the case of GPS, the refresh cycle is every 2 h; for Galileo, it is shortened to every 20 min. The update frequency for the navigation message depends on the rate at which the quality of the chosen orbit representation deteriorates over time compared with the expected positioning accuracy provided to users. For lunar systems, it is crucial to consider the visibility of satellites from the ground segment, which is critical for data transmission.
6.3 Trade-off for the Representation of Broadcast Orbits
Keplerian parameters with corrections are a natural choice for representing the broadcast ephemerides for a navigation system designed for the Moon, as they require a smaller number of bits when compared with Chebyshev representation. The eccentric orbit design, however, introduces challenges in modeling the orbit representation and in representing the ephemerides. The Keplerian elements must be converted into Cartesian coordinates for orbit extrapolation and for introducing empirical accelerations in the radial, along-track, and cross-track directions; thus, the receiver must be capable of conducting numerical integration, which introduces additional computation complexity.
For orbit representation based on Chebyshev polynomials, the accuracy of the function fit increases with the degree of the polynomial. Hence, the user does not need to receive a full message to obtain a satellite position. However, as the degree of the polynomial received by the user increases, the orbit representation obtained by the user becomes more accurate. Moreover, the transformation parameters between the lunar and celestial reference frames can also be based on Chebyshev polynomials, simplifying the algorithms that must be applied and implemented. Receiving only the first five coefficients of the Chebyshev polynomial provides an orbit representation accuracy of roughly 30 cm in 50% of windows, which is sufficient for most real-time navigation purposes on the Moon.
7 LUNAR REFERENCE FRAME AND TIME SCALE
The definition of a lunar reference system and its precise realization are essential for the LRNS, in the same way that the International Terrestrial Reference Frame is the basis of terrestrial GNSSs. The time scale adopted by users in the lunar environment is equally important. In this section, after a brief review of existing reference frames and proposed time scales, we suggest the adoption of a reference frame and time scale that suits the needs of most lunar users by broadcasting a navigation message built on that frame and time scale.
At present, two slightly different reference systems based on the Moon are commonly used to define a lunar body-fixed coordinate system: the mean Earth (ME)/rotation axis (or polar axis) reference system and the principal axis (PA) reference system (Archinal et al., 2015; Williams & Boggs, 2008). The former was used at the beginning of lunar observation and is commonly adopted for archiving and data distribution purposes related to lunar surface or topography. The latter corresponds to an orientation in which the lunar tensor of inertia is diagonal. Currently, with our highly accurate knowledge of the lunar gravity field, the PA system is an attractive alternative to the old ME system. In addition, the dynamically defined PA system is more frequently applied to bodies with thick atmospheres, such as Venus and Titan, therefore easing the adoption of this system across the solar system. Even if cartographic products are referenced to the ME system, the more modern and accurate maps are in a digital format, and re-referencing these products to the PA system would be a simple matter. More effort would certainly be needed for older analog maps, although advanced tools are currently available to scan and digitize analog maps; cartographic coordinates could then be easily applied. It is conceivable that both systems will coexist for some time, in spite of the advantages of the PA frame. In any case, a transformation between the two reference systems is possible by means of three static Euler angles that depend on the lunar gravity field coefficients and dissipative models. Thus, a lunar ephemeris is necessary to carry out the transformation.
7.1 Inertial and Kinematically Non-Rotating systems
Moon-centered inertial reference systems are useful for OD of spacecraft in the cislunar environment. Current definitions of inertial systems are based on the International Astronomical Union (IAU) B3 recommendations (Soffel et al., 2003; Petit and Luzum, 2010) and relativistic descriptions of the barycentric celestial reference system (BCRS) and geocentric celestial reference system (GCRS) reported by Damour et al. (1991). The BCRS metric is taken to be kinematically fixed with respect to distant quasars, as defined by Soffel et al. (2003) and following harmonic gauge conditions. A kinematically non-rotating reference system centered on the Earth’s center of mass has been defined in the same type of framework, leading to the definition of the GCRS. Time scales derived from the BCRS and the GCRS metric have been reported by Soffel et al. (2003).
For the Moon, Turyshev et al. (2013) proposed to adapt the GCRS definition to the Moon in the context of the GRAIL mission (recall, however, that the International Astronomical Union is in charge of defining the LCRS). A description of such a Moon-centered inertial system is given in Section 6.6, with a definition of the Moon-centered metric and time scale in Section 6.7. In Section 6.8, an assessment of the accuracy of two possible realizations of a lunar celestial reference system (LCRS) is given by comparing results obtained with DE421 (Folkner et al., 2014) and INPOP19a (Fienga et al., 2019).
7.2 Principal Axis System
The PA system is defined by the PA orientation of the inertia matrix of the Moon with constant tide contributions. Indeed, this is a useful reference system for writing and integrating the rotational equation. Its materialization is realized from a temporal series of the Moon orientation, coming from ephemerides, and described through the rotation Euler angles (ψ: precession angle, θ: nutation angle, φ: proper rotation). The rotational angles are extracted from the lunar ephemerides (Jet Propulsion Lab [JPL] development ephemerides [DE] or Intégrateur Numérique Planétaire de l’Observatoire de Paris [INPOP] ephemerides produced by the Institut de Mécanique Céleste et de Calcul des Ephémérides and Observatoire de la Cote d’Azur). These two ephemerides are similar in terms of weighted root mean square residuals. Let I be a vector in the lunar celestial reference frame (LCRS) and let PA be a vector in the PA reference frame, with Rx and Rz as elementary rotational matrices. The direction and transformation are written as follows (Folkner et al., 2014):
12
7.3 Mean Earth Direction and Rotation System (EPM)
Because the Moon is not a purely triaxial ellipsoid, the average orientation of the PA does not coincide with the ME direction. Figure 10 shows the position of the Earth in the lunar PA reference system over 1 and 6 years. The loops indicate the geometrical libration of the Moon. The average of the loops is not exactly zero, and there is a constant offset. The ME/polar axis reference system is defined by the z-axis as the mean rotational pole. The prime meridian (0° longitude) is defined by the ME direction. The intersection of the lunar equator and prime meridian occurs at what can be called the Moon’s “mean sub-Earth point.” This system is an idealization; a practical attempt to determine these mean directions with high accuracy would depend on the approach and time interval used. This reference system is defined from the numerical ephemerides of the PA system through the following rotation matrix relationship (Folkner et al., 2014):
Direction from the Moon to Earth over 1 year (left) and 6 years (right)
u1, u2, and u3 are cosines of the direction vector of the Earth as seen in the PA reference frame of the Moon (no units).
13
where Ic is the Moon’s moment of inertia, p1c, p2c, are constant coordinates of the ecliptic pole in the PA reference frame, and τc and σc are constant libration angles. The offset between these coordinate systems for a point on the lunar surface is approximately 860 m and depends on the ephemeris and the associated gravitational field. The inverse transformation is as follows:
14
7.4 Lunar Laser Ranging Retroreflector as Lunar Point Control
There are five points on the Moon whose positions are very accurately measured. These points correspond to lunar laser ranging retroreflectors (LLRRs), distributed over the equatorial and mid-latitude regions of the Moon. These LLRRs can be used as control points of a lunar geodetic network, given their high stability over time. The stability of the LLRR positions has been recently discussed by Williams and Boggs (2021), who provided the coordinate rates of the LLRRs. The weighted average rate of the four LLRs is 8.3 ± 3.1 mm/year, corresponding to 20 cm over 25 years. The origin of this shift is not known, but the authors argued that it could be due to mismodeling of rotational motion. Finally, the formal uncertainties in the LLRR positions are approximately 3 cm in the INPOP ephemerides, and the global uncertainty in position is 40 cm for the JPL DE (Williams & Boggs, 2008). These uncertainties of few tens of centimeters represent the internal precision in ephemerides.
A comparison of various ephemerides provides their external accuracy, corresponding to approximately 2 m for EPM (Pavlov, 2020) compared with the DE (Park et al., 2021; Folkner et al., 2014) and INPOP (Viswanathan et al., 2018; Fienga et al., 2019; Fienga et al., 2021) values and 1 m for INPOP and DE440. Besides comparing LLRR coordinates obtained in different PA frames, one can also compare these coordinates with those obtained via other methods, such as altimetry or imaging.
7.5 Proposition of Lunar Reference Systems: A Moon-Centered System (LCRS)
Based on the concept of the GCRS given in the IAU 2000 Resolution B1.3 (Soffel et al., 2003; Petit & Luzum, 2010), we propose to define the LCRS, centered at the Moon’s center of mass, associated with this metric tensor based on the work of Damour et al. (1991), but adapted to the gravitational environment of the Moon. A lunar time scale is also defined. This system will be used for navigation and for integrating the equations of motion at the vicinity of the Moon.
By analogy with the definition of the GCRS metric tensor and the IAU 2000 Resolution B1.3 (Soffel et al., 2003; Petit and Luzum, 2010), we consider the selenocentric metric tensor Sab with selenocentric coordinates (T, X), where T is the lunar coordinate time (TCL) and X is the coordinate vector of an orbiter in the LCRS. Similarly, in the BCRS, t and x denote the barycentric time scale (barycentric coordinate time [TCB] or barycentric dynamical time [TDB]) and the coordinate vector of an orbiter, respectively. The selenocentric tensor is in the same form as the barycentric or geocentric tensor but with Moon-centered potentials L(T, X) and La(T, X) as follows:
15
The potentials L and La should be split into two parts: potentials LS and arising from the gravitational action of the Moon and external terms Lext and La due to tidal and inertial effects. The external parts of the metric potentials are assumed to vanish at the selenocenter and admit an expansion into positive powers of XL. Explicitly, we have the following:
16
17
where AM is the Moon’s angular momentum. LS corresponds to the lunar Newtonian gravitational potential, and Lext is the tidal contribution produced by all solar system bodies (excluding the Moon) estimated at the center of mass of the Moon (origin of the LCRS) at T = TCL. The expression of the potentials Lext, LS, and La can be found, for example, in the work by Turyshev et al. (2013).
Considering T = TCL, t = TCB, rM = x − xM, i.e., the differences between the vectors of the BCRS barycentric position of the orbiter x and the barycentric position of the Moon xM, with vM and aM being the vectors of its barycentric velocity and acceleration, the transformation between the BCRS and LCRS is given explicitly. The expression of the LCRS coordinate vector X from the BCRS x coordinate vector (appearing in the vector rM = x − xM) is given by the following:
18
Here, lext (XM) = ∑A≠M GMa/rAM + O(c−2), and rAM = xA − xM, the Moon-centered vector of body A in the BCRS.
The time transformation between BCRS time t and LCRS time T is as follows:
19
where , and are the coordinates of the vectors vM, rM, and aM, respectively, with the following:
20
7.6 Proposition of a Lunar Time Scale for the Moon-Centered System (LCRS)
The UTC time scale is used for time-tagging observations from the ground. For the definition of selenodetic frames or when one uses Earth-based observations for studying the dynamics of the Moon, the UTC time scale is transformed into TDB (Fienga et al., 2009). The GPS time can be used as an intermediate time scale between lunar orbiters and ground-based stations, a natural choice for receivers of Earth GNSS signals in lunar orbit. A recent paper by Ashby and Patla (2024) established a framework for the estimation of clock rates on the Moon.
For time-tagging on the Moon surface or for a constellation orbiting the Moon, we define the TCL, the relativistic time scale at the center of mass of the Moon, similar to the geocentric coordinate time (TCG), the relativistic time scale at the geocenter (Soffel et al., 2003; Petit & Luzum, 2010). Based on the previous equations, we can define TCL as related to TCB in the same way that TCG is related to TCB (Damour et al., 1991):
21
with:
22
and:
23
Subscripts A and B denote massive bodies, L corresponds to the Moon, MA is the mass of body A, rLA = xL − xA (vectorial differences), rLA is the norm of the vector rLA, xA is the barycentric vector of position of the massive body A, and vA and aA are the barycentric vectors of velocity and acceleration of the massive body A in TCB. αL is approximately 1.48 × 10−8 in the work by Turyshev et al. (2013), and as explained by Nelson and Ely (2006) and Turyshev et al. (2013), the βL term (Equation (21)) is negligible for the present applications.
In the same way that TCL can be defined relative to TCB (Equation (21)), it is also possible to describe TCL relative to TCG in the GCRS. In this case, we have the following:
24
where αL/G and βL/G (which is negligible and whose expression is not reported) have the same definitions as in Equations (22) and (23) but with geocentric positions and velocities:
25
In contrast, terrestrial time (TT) is defined relative to TCG as follows (Petit & Luzum, 2010):
26
with LG = 6.969290134 10−10. By considering that TCL − TT = (TCL − TCG) + (TCG − TT), with Equation (21) and the current definition of TT (Equation (26)), we can derive the following:
27
Explicit expressions for the transformations between local lunar-centered reference frame coordinates of an object in the vicinity of the Moon, X(TCL), and the coordinates of the same object in the BCRS relative to the Moon, rM (TDB) (shown as rM as in Equation (18)) have been reported by Soffel et al. (2003) and Turyshev et al. (2013).
7.7 Assessment of Realization Accuracy
The LCRS can be realized using different lunar and planetary ephemerides. Here, we consider two sets of ephemerides, namely, DE421 (Williams et al., 2022), which is the present reference for the definition of the lunar ME frame (see Section 6.3), and INPOP19a (Fienga et al., 2019). These two ephemerides differ, as the DE421 model does not account for core–mantle interactions or the shape of the fluid core. More than 11 years of additional observations were also included in the INPOP19a construction. The differences in terms of localization of the Moon relative to the Earth using DE421 and INPOP19a over 5 years are plotted in Figure 11. The maximum difference of approximately 1.5 m is obtained in the Z direction. The ratio between the gravitational mass of the Earth–Moon barycenter obtained with DE421 and with INPOP19a differs from unity by (1.2 ± 10−11) × 10−8.
Differences (in meters) in Moon barycenter positions relative to the Earth estimated with DE421 and INPOP19a in the International Celestial Reference Frame (ICRF)
8 FINAL REMARKS
The proposed architecture of a lunar radio navigation system, consisting of four spacecraft in ELFOs, was developed based on four key guidelines: simplification of the ground network, utilization of high-technology-readiness-level hardware, cost-effective scalability to a larger constellation, and excellent performance in OD and time transfer. The adoption of an MSPA configuration for spacecraft tracking is key to the first goal, as it entails the use of small antennas (less than 30 cm in diameter), which facilitates the deployment of network terminals and reduces the overall cost of the ground infrastructure. Furthermore, the small antenna size enables the design and assembly of compact and easily transportable terminals, enhancing the network’s flexibility. To optimize cost and ease of deployment, it is advantageous for the LRNS backbone ground infrastructure to be co-located with existing deep space tracking sites or radio astronomical antennas. These ground antennas already possess hydrogen maser frequency standards and fast communication lines, which are crucial to the overall system.
A drawback of the proposed configuration is the requirement for small (30-cm-diameter) steerable antennas onboard the spacecraft. However, steerable antennas provide sufficient data rates for telecommands and telemetry, supporting the frequent upload of navigation messages and the download of housekeeping data and code epoch time stamps. Owing to nearly continuous visibility, the daily data volume achieved with the proposed architecture is comparable to that obtained with a much larger antenna (e.g., 10–15 m in diameter) tracking the constellation’s nodes sequentially. This result further emphasizes the efficiency and practicality of the system design.
Most elements of the radio system have a high technology readiness level (>5), despite the selection of the K-band for the radio link. A notable exception is the CDMA back-end receiver, which is currently being developed at the breadboard level. This receiver is specifically designed to support SS signals with a bandwidth of 6 MHz. It is worth mentioning that CDMA/CDM-M modulations have not yet been used in deep space communications. Nonetheless, existing digital transponders can be readily modified to meet the requirements of the proposed radio link configuration. In terms of passive elements, there are already commercially available, space-qualified, steerable antennas operating at the K-band. These off-the-shelf antennas are fully compatible with an accommodation on small-sats and can be used effectively within the system.
Although the constellation considered for ESA’s Moonlight initiative is designed to support missions in the southern polar regions of the Moon, the future of lunar exploration will certainly need extended coverage and an increased number of spacecraft. The proposed architecture would make such an expansion straightforward. The ground support to an extended constellation requires only the deployment of additional ground terminals, each supporting a separate set of nodes. The selection of the K-band as the operating frequency is a key factor in enabling this expandability. Compared with the X-band, the K-band offers a significantly larger available bandwidth (500 MHz compared with 50 MHz), providing more flexibility in designing the signal structure. The use of SS signals at large chip rates (24–25 Mcps in our concept), essential for attaining good ranging accuracies (33 cm for a 10-s integration time), is possible only in the K-band. By comparison, the recommended channel bandwidths at the X-band (6 MHz) would allow range measurements with a thermal jitter of approximately 1.3 m. Furthermore, the K-band exhibits improved immunity to charged particle effects in radiometric measurements, making it advantageous for accurate and precise data acquisition. Therefore, by adopting the K-band, the proposed architecture not only allows for straightforward expansion but also enhances the system’s performance in terms of ranging accuracy and immunity to charged particle effects.
The proposed architecture, utilizing K-band and high-chip-rate SS signals, offers several advantages, including high-precision and bias-free range and range-rate measurements. Moreover, the adoption of the MSPA tracking configuration makes a third data type (SBI) available for precise OD. Implementing SBI is straightforward, as it involves differencing the receiver phases from any pair of spacecraft without requiring additional hardware. The OD accuracy relies heavily on the quality of radiometric measurements. Numerical simulations conducted on the proposed architecture demonstrate that the SISE consistently meets the Moonlight initiative’s requirements. Additionally, the simulations show that the ephemerides aging is gradual enough to only necessitate uploading a new navigation message at intervals of several hours, ensuring reliable and up-to-date orbital data to the end user.
The architecture for the LRNS is augmented by a proposal for an efficient and accurate navigation message to be broadcasted to lunar users. Our comparative analysis indicates that, for the highly elliptic orbits considered, Chebyshev polynomials are the preferred choice. Not only are Chebyshev polynomials used for representing planetary ephemerides, but, should an increased accuracy be needed in the representation of the constellation orbit, augmenting the navigation message by additional Chebyshev coefficients would provide a simple and straightforward solution.
Finally, we proposed lunar reference frames (both rotating and non-rotating) and a lunar time scale that would be easily integrated in the LRNS and adopted by lunar users. The metric underlying these frames and time scale is largely based on previous work, with the concurrence of the IAU. Although the ultimate choice for reference frames and time scales will be determined by space agencies, the IAU, and BIPM, we offer our contribution to ongoing and future discussions.
HOW TO CITE THIS ARTICLE
Iess, L., Di Benedetto, M., Boscagli, G., Racioppa, P., Sesta, A., De Marchi, F., Cappuccio, P., Durante, D., Molli, S., Plumaris, M. K., Tartaglia, P., Fienga, A., Rambeaux, N., Santi, F., Pastina, D., Linty, N., Sosnica, K., Bury, G., Zajdel, R., ...Ventura-Traveset, J. (2025). A novel orbit determination and time synchronization architecture for a radio navigation satellite constellation in the cislunar environment. NAVIGATION, 72(3). https://doi.org/10.33012/navi.713
ACKNOWLEDGMENTS
This work was carried in the context of the ESA study “Fundamental techniques, models and algorithms for a lunar radio navigation system,” contract no. 4000136075/21/NL/CRS.
This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.
REFERENCES
- ↵Amin, C. (2016). Cluster multiple spacecraft per aperture operations. Proc. of the 14th International Conference on Space Operations (SpaceOps 2016), Daejeon, Korea, 2016–2639. https://doi.org/10.2514/6.2016-2639
- ↵Ashby, N., & Patla, B. R. (2024). A relativistic framework to estimate clock rates on the Moon. The Astronomical Journal, 168(3), 112. https://doi.org/10.3847/1538-3881/ad643a
- ↵Asmar, S. W., Armstrong, J. W., Iess, L., & Tortora, P. (2005). Spacecraft Doppler tracking: Noise budget and accuracy achievable in precision radio science observations. Radio Science, 40(2), 1–9. https://doi.org/10.1029/2004rs003101
- ↵Barnes, J. A., Andrew, R., Cutler, L., Healey, D., Leeson, D., McGunigal, T. E., Mullen, J. A., Smith, W. L., Sydnor, R. L., Vessot, R. F. C., & Winkler, G. M. R. (1971). Characterization of frequency stability. IEEE Transactions on Instrumentation and Measurement, IM–20(2), 105–120. https://doi.org/10.1109/tim.1971.5570702
- ↵Bauer, S., Hussmann, H., Oberst, J., Dirkx, D., Mao, D., Neumann, G. A., Mazarico, E., Torrence, M. H., McGarry, J. F., Smith, D. E., & Zuber, M. T. (2017). Analysis of one-way laser ranging data to LRO, time transfer and clock characterization. Icarus, 283, 38–54. https://doi.org/10.1016/j.icarus.2016.09.026
- ↵Bender, P. L. (1994). Proposed microwave transponders for early lunar robotic landers. Advances in Space Research, 14(6), 233–242. https://doi.org/10.1016/0273-1177(94)90033-7
- ↵BIPM (2023). Circular T 425, ISSN 1143-1393. Retrieved from https://www.bipm.org/en/time-ftp/circular-t
- ↵Burt, E. A., Prestage, J. D., Tjoelker, R. L., Enzer, D. G., Kuang, D., Murphy, D. W., Robison, D. E., Seubert, J. M., Wang, R. T., & Ely, T. A. (2021). Demonstration of a trapped-ion atomic clock in space. Nature, 595(7865), 43–47. https://doi.org/10.1038/s41586-021-03571-7
- ↵Campa, C., Gianfelici, D., Abelló, R., & Martí, S. (2022). Multiple spacecraft per aperture with code division multiple access modulation. Proc. of the 9th ESA International Workshop on Tracking, Telemetry, and Command Systems for Space Applications (TT&C 2022 Workshop ESA/ESTEC). https://www.researchgate.net/publication/370609327_TTC_2022_MULTIPLE_SPACECRAFT_PER_APERTURE_WITH_CODE_DIVISION_MULTIPLE_ACCESS_MODULATION
- ↵Cappuccio, P., Hickey, A., Durante, D., Di Benedetto, M., Iess, L., De Marchi, F., Plainaki, C., Milillo, A., & Mura, A. (2020). Ganymede’s gravity, tides and rotational state from JUICE’s 3GM experiment simulation. Planetary and Space Science, 187, 104902. https://doi.org/10.1016/j.pss.2020.104902
- ↵Cappuccio, P., Notaro, V., Di Ruscio, A., Iess, L., Genova, A., Durante, D., Di Stefano, I., Asmar, S. W., Ciarcia, S., & Simone, L. (2020). Report on first inflight data of BepiColombo’s Mercury Orbiter radio science experiment. IEEE Transactions on Aerospace and Electronic Systems, 56(6), 4984–4988. https://doi.org/10.1109/taes.2020.3008577
- ↵Consultative Committee for Space Data Systems. (2011). CCSDS 415.1-B-1: Data transmission and PN ranging for 2 GHz CDMA (Blue Book). https://ccsds.org/Pubs/415x1b1.pdf
- ↵Consultative Committee for Space Data Systems. (2013). CSDS 415.0-G-1: Data transmission and PN ranging for 2 GHz CDMA (Green Book). https://ccsds.org/Pubs/415x0g1.pdf
- ↵Consultative Committee for Space Data Systems. (2014). Pseudo-noise (PN) ranging systems. Report concerning space data system standards (Green Book). Issue 2, CCSDS 414.0-G-2. https://ccsds.org/Pubs/414x0g2.pdf
- ↵Consultative Committee for Space Data Systems (CCSDS). (2019). Delta-DOR technical characteristics and performance (Green Book). Issue 2, CCSDS 500.1-G-2. https://ccsds.org/Pubs/500x1g2.pdf
- ↵Consultative Committee for Space Data Systems (CCSDS). (2021). Radio frequency and modulation system (Blue Book). Issue 32, CCSDS 401.0-B-32 https://ccsds.org/Pubs/401x0b32.pdf
- ↵Damour, T., Soffel, M., & Xu, C. (1991). General-relativistic celestial mechanics. I. Method and definition of reference systems. Physical Review D: Particles and Fields, 43(10), 3273–3307. https://doi.org/10.1103/physrevd.43.3273
- ↵Delva, P., Meynadier, F., Poncin-Lafitte, C. L., Laurent, P., & Wolf, P. (2012). Time and frequency transfer with a microwave link in the ACES/PHARAO mission. Proc. of the 2012 European Frequency and Time Forum, Gothenburg, Sweden, 28–35. https://doi.org/10.1109/eftf.2012.6502327
- ↵Donà, G., & Iess, L. (2017). CDMA implementation for TT&C and precision navigation - Scenario analysis and signal definition. [Technical note TN-002-1]. Thales Alenia Space Italia, ESA contract no. 4000117076/16/NL/FE.
- ↵Donà, G., & Rovelli, D. (2016). Code division multiplexing transmitter options for single aperture multiple link applications. Proc. of the Third ESA Workshop on Advanced Flexible Telecom Payloads (ESA/ESTEC), Noordwijk, Netherlands.
- ↵DSMS Telecommunications Link Design Handbook. (2015). Atmospheric and environmental effects. DSN Doc. 810-005, 105, Rev. E. https://deepspace.jpl.nasa.gov/dsndocs/810-005/105/105E.pdf
- ↵Duchayne, L., Mercier, F., & Wolf, P. (2009). Orbit determination for next generation space clocks. Astronomy and Astrophysics, 504(2), 653–661. https://doi.org/10.1051/0004-6361/200809613
- ↵European Space Agency (ESA). (2016). Multiple spacecraft per aperture with code division multiple access modulation and linearizer [Statement of work]. ESA-GRST-STU-SOW-0010. ESA-STAR log-in (granted upon request) is required to download the statement of work. https://esastar-publication-ext.sso.esa.int/ESATenderActions/details/7257
- ↵Fienga, A., Deram, P., Di Ruscio, A., Viswanathan, V., Camargo, J., Bernus, L., Gastineau, M., & Laskar, J. (2021). INPOP21a planetary ephemerides. Notes Scientifiques et Techniques de l’Institut de Mecanique Celeste, 110, 1–19. https://www.imcce.fr/content/medias/recherche/equipes/asd/inpop/inpop21a.pdf
- ↵Fienga, A., Deram, P., Viswanathan, V., Di Ruscio, A., Bernus, L., Durante, D., Gastineau, M., & Laskar, J. (2019). INPOP19a planetary ephemerides. Notes Scientifiques et Techniques de l’Institut de Mecanique Celeste, 109, 1–35. https://www.imcce.fr/content/medias/recherche/equipes/asd/inpop/inpop19a_20191214.pdf
- ↵Fienga, A., Laskar, J., Morley, T., Manche, H., Kuchynka, P., Le Poncin-Lafitte, C., Budnik, F., Gastineau, M., & Somenzi, L. (2009). INPOP08, a 4-D planetary ephemeris: From asteroid and time-scale computations to ESA Mars Express and Venus Express contributions. Astronomy and Astrophysics, 507, 1675–1686. https://doi.org/10.1051/0004-6361/200911755
- ↵Folkner, W. M., Williams, J. G., Boggs, D. H., Park, R. S., & Kuchynka, P. (2014). The planetary and lunar ephemerides DE430 and DE431. Interplanetary Network Progress Report, 196(1), 42–196. https://ipnpr.jpl.nasa.gov/progress_report/42-196/196C.pdf
- ↵Gelfusa, D., Cialfi, F., Salerno, N., & Simone, L. (2019). Integrated deep space & radioscience TT&C transponder (IDST). Proc. of the 8th International Workshop on Tracking, Telemetry, and Command for Space, Darmstadt, Germany. https://atpi.eventsair.com/QuickEventWebsitePortal/19a06-8th-ttc-2019/website/Agenda/AgendaItemDetail?id=017c5865-5958-495c-ba53-458f1ea5356c
- ↵Giordano, P., Grenier, A., Zoccarato, P., Bucci, L., Cropp, A., Swinden, R., Otero, D., Eldali, W., Care, W., Duvet, L., Hufenbach, B., Joly, F., & Ventura-Traveset, J. (2021). Moonlight navigation service - how to land on peaks of eternal light. Proc. of the 72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25–29. https://www.researchgate.net/publication/358266257_Moonlight_navigation_service_-_how_to_land_on_peaks_of_eternal_light
- ↵Gregnanin, M., Bertotti, B., Chersich, M., Fermi, M., Iess, L., Simone, L., Tortora, P., & Williams, J. G. (2012). Same beam interferometry as a tool for the investigation of the lunar interior. Planetary and Space Science, 74(1), 194–201. https://doi.org/10.1016/j.pss.2012.08.027
- ↵Gregnanin, M., Yseboodt, M., Dehant, V., Iess, L., & Van Hoolst, T. (2014). Estimation of Mars geophysical information through same beam interferometry. Proc. of the European Planetary Science Congress EPSC 2014, Cascais, Portugal, Vol. 9. https://meetingorganizer.copernicus.org/EPSC2014/EPSC2014-395.pdf
- ↵Holmes, J. K. (1982). Coherent spread spectrum systems. Wiley-Interscience. https://ui.adsabs.harvard.edu/abs/1982wi...bookR....H/abstract
- ↵Howe, D. A., Hanson, D. W., Jesperscn, J. L., & Lombardi, M. A. (1989). Satellite two-way time transfer: Fundamentals and recent progress. Proc. of the 21th Annual Precise Time and Time Interval Systems and Applications Meeting, Redondo Beach, CA, 117–130. https://www.ion.org/publications/abstract.cfm?articleID=14510
- ↵Iess, L., Di Benedetto, M., James, N., Mercolino, M., Simone, L., & Tortora, P. (2014). Astra: Interdisciplinary study on enhancement of the end-to-end accuracy for spacecraft tracking techniques. Acta Astronautica, 94(2), 699–707. https://doi.org/10.1016/j.actaastro.2013.06.011
- ↵Iess, L., Folkner, W. M., Durante, D., Parisi, M., Kaspi, Y., Galanti, E., Guillot, T., Hubbard, W. B., Stevenson, D. J., Anderson, J. D., Buccino, D., Casajus, L. G., Milani, A., Park, R., Racioppa, P., Serra, D., Tortora, P., Zannoni, M., Cao, H., . . . Bolton, S. J. (2018). Measurement of Jupiter’s asymmetric gravity field. Nature, 555(7695), 220–222. https://doi.org/10.1038/nature25776
- ↵Iess, L., Asmar, S. W., Cappuccio, P., Cascioli, G., De Marchi, F., di Stefano, I., Genova, A., Ashby, N., Barriot, J. P., Bender, P., Benedetto, C., Border, J. S., Budnik, F., Ciarcia, S., Damour, T., Dehant, V., Di Achille, G., Di Ruscio, A., Fienga, A, ...Wu, X. (2021). Gravity, geodesy and fundamental physics with BepiColombo’s MORE investigation. Space Science Reviews, 217(1), 21. https://doi.org/10.1007/s11214-021-00800-3
- ↵Iess, L., Di Benedetto, M., Boscagli, G., Racioppa, P., Sesta, A., De Marchi, F., Cappuccio, P., Durante, D., Molli, S., Plumaris, M. K., Tartaglia, P., Santi, F., Fienga, A., Linty, N., Sośnica, K., Belfi, J., Giordano, P., Swinden, R., & Ventura-Traveset, J. (2023). High performance orbit determination and time synchronization for lunar radio navigation systems. Proc. of the 36th International Technical Meeting of the Satellite Division of the Institute of Navigation (ION GNSS+ 2023), Denver, CO, 4029–4050. https://doi.org/10.33012/2023.19428
- ↵ITU-R TF.1153-4. (2015). The operational use of two-way satellite and frequency transfer employing pseudorandom noise codes’ time signals and frequency standards emissions. Retrieved from https://www.itu.int/dms_pubrec/itu-r/rec/tf/R-REC-TF.1153-4-201508-I!!PDF-E.pdf
- ↵Levine, J. (2008). A review of time and frequency transfer methods. Metrologia, 45(6), S162. https://doi.org/10.1088/0026-1394/45/6/S22
- ↵Melman, F. T., Zoccarato, P., Orgel, C., Swinden, R., Giordano, P., & Ventura-Traveset, J. (2022). LCNS positioning of a lunar surface rover using a DEM-based altitude constraint. Remote Sensing, 14(16), 3942. https://doi.org/10.3390/rs14163942
- ↵Meurer, M., & Antreich, F. (2017). Signals and modulation. In P. J. G. Teunissen & O. Montenbruck (Eds.), Springer handbook of global navigation satellite systems (pp. 91–119). https://doi.org/10.1007/978-3-319-42928-1_4
- ↵Moyer, T. D. (2003). Formulation for observed and computed values of deep space network data types for navigation. John Wiley and Sons. https://doi.org/10.1002/0471728470
- ↵Nelson, R., & Ely, T. (2006). Relativistic transformations for time synchronization and dissemination in the solar system. Proc. of the 38th Annual Precise Time and Time Interval Systems and Applications Meeting, Reston, VA, 305–317. https://www.ion.org/publications/abstract.cfm?articleID=13655
- ↵Park, R. S., Folkner, W. M., Williams, J. G., & Boggs, D. H. (2021). The JPL planetary and lunar ephemerides DE440 and DE441. The Astronomical Journal, 161(3), 105. https://doi.org.10.3847/1538-3881/abd414
- ↵Pavlov, D. (2020). Role of lunar laser ranging in realization of terrestrial, lunar, and ephemeris reference frames. Journal of Geodesy, 94(1), 5. https://doi.org/10.1007/s00190-019-01333-y
- ↵Petit, G., & Luzum, B. (2010). IERS conventions (2010). https://apps.dtic.mil/sti/tr/pdf/ADA535671.pdf
- ↵Sank, V.J. (1993). Using the USCCS for sub microsecond spacecraft clock calibration. Proc. of the NASA Goddard Space Flight Center Networks Technology Conference, Greenbelt, MD. https://ntrs.nasa.gov/api/citations/19940016873/downloads/19940016873.pdf
- ↵Schutz, B., Tapley, B., & Born, G. H. (2004). Statistical orbit determination. Elsevier. https://doi.org/10.1016/B978-0-12-683630-1.X5019-X
- ↵Sesta, A., Durante, D., Boscagli, G., Cappuccio, P., Di Benedetto, M., di Stefano, I., Plumaris, M. K., Racioppa, P., Iess, L., Giordano, P., Swinden, R., & Ventura-Traveset, J. (2025). ATLAS: Orbit determination and time transfer for a lunar radio navigation system. NAVIGATION, 72(2). https://doi.org/10.33012/navi.701
- ↵Simone, L., Fittipaldi, G., Barigelli, A., Guerrucci, R. (2010). Techniques to increase the secure TT&C anti jamming capabilities beyond the processing gain. In Proceedings of TTC 2010 5th ESA International Workshop on Tracking, Telemetry and Command Systems for Space Applications.
- ↵Soffel, M., Klioner, S. A., Petit, G., Wolf, P., Kopeikin, S. M., Bretagnon, P., Brumberg, V. A., Capitaine, N., Damour, T., Fukushima, T., Guinot, B., Huang, T.-Y., Lindegren, L., Ma, C., Nordtvedt, K., Ries, J. C., Seidelmann, P. K., Vokrouhlický, D., Will, C. M., & Xu, C. (2003). The IAU 2000 resolutions for astrometry, celestial mechanics, and metrology in the relativistic framework: Explanatory supplement. The Astronomical Journal, 126(6), 2687. https://doi.org/10.1086/378162
- ↵Steigenberger, P., & Montenbruck, O. (2017). Galileo status: Orbits, clocks, and positioning. GPS Solutions, 21(2), 319–331. https://doi.org/10.1007/s10291-016-0566-5
- ↵Thornton, C. L., & Border, J. S. (2003). Radiometric tracking techniques for deep-space navigation. John Wiley & Sons. https://descanso.jpl.nasa.gov/monograph/series1/Descanso1_all.pdf
- ↵Turyshev, S. G., Toth, V. T., & Sazhin, M. V. (2013). General relativistic observables of the GRAIL mission. Physical Review D, 87(2), 024020. https://doi.org/10.1103/PhysRevD.87.024020
- ↵Viswanathan, V., Fienga, A., Minazzoli, O., Bernus, L., Laskar, J., & Gastineau, M. (2018). The new lunar ephemeris INPOP17a and its application to fundamental physics. Monthly Notices of the Royal Astronomical Society, 476(2), 1877–1888. https://doi.org/10.1093/mnras/sty096
- ↵Williams, J. G., & Boggs, D. H. (2008). DE421: Lunar orbit, physical librations, and surface coordinates [Technical report]. JPL Interoffice Memorandum IOM 335-JW. ftp://ssd.jpl.nasa.gov/pub/eph/planets/ioms/de421_moon_coord_iom.pdf
- ↵Williams, J. G., & Boggs, D. H. (2021). Testing the motion of lunar retroreflectors. Journal of Geophysical Research: Planets, 126(11), e2021JE006920. https://doi.org/10.1029/2021je006920
- ↵Williams, J. G., Boggs, D. H., & Currie, D. G. (2022). Next-generation laser ranging at lunar geophysical network and commercial lander payload service sites. The Planetary Science Journal, 3(6), 136. https://doi.org/10.3847/PSJ/ac6c25












