This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

The following article is Open access

The JPL Planetary and Lunar Ephemerides DE440 and DE441

, , , and

Published 2021 February 8 © 2021 The Author(s). Published by The American Astronomical Society.
, , Citation Ryan S. Park et al 2021 AJ 161 105 DOI 10.3847/1538-3881/abd414

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1538-3881/161/3/105

Abstract

The planetary and lunar ephemerides called DE440 and DE441 have been generated by fitting numerically integrated orbits to ground-based and space-based observations. Compared to the previous general-purpose ephemerides DE430, seven years of new data have been added to compute DE440 and DE441, with improved dynamical models and data calibration. The orbit of Jupiter has improved substantially by fitting to the Juno radio range and Very Long Baseline Array (VLBA) data of the Juno spacecraft. The orbit of Saturn has been improved by radio range and VLBA data of the Cassini spacecraft, with improved estimation of the spacecraft orbit. The orbit of Pluto has been improved from use of stellar occultation data reduced against the Gaia star catalog. The ephemerides DE440 and DE441 are fit to the same data set, but DE441 assumes no damping between the lunar liquid core and the solid mantle, which avoids a divergence when integrated backward in time. Therefore, DE441 is less accurate than DE440 for the current century, but covers a much longer duration of years −13,200 to +17,191, compared to DE440 covering years 1550–2650.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Modern-day planetary ephemerides are computed by fitting numerically integrated orbits to various types of ground-based and space-based observations (Folkner et al. 2014; Pitjeva & Pitjev 2018; Fienga et al. 2020). The Jet Propulsion Laboratory's (JPL) planetary and lunar ephemerides Development Ephemeris (DE) series includes the positions of the Sun, the barycenters of eight planetary systems, the Moon, the Pluto system barycenter, and lunar libration angles, as well as their associated velocities. The high-precision orbits and lunar rotations around the three axes have a wide range of practical and fundamental applications (Thornton & Border 2003; Park et al. 2020b; Vallisneri et al. 2020; U.S. Nautical Almanac Office & Her Majesty's Nautical Almanac Office 2018). Without an update, however, the errors in orbits grow for several reasons. For Jupiter (Juno 2016–2020 < 12 yr period), Saturn (Cassini 2004–2018, <30 yr period), Uranus, Neptune, and Pluto the high-quality data are available for less than one orbit. For Mars, the range data is of high quality, but the main-belt asteroid masses are a limitation partly due to the large number of asteroids and partly due to long-period perturbations (Folkner et al. 2014). In the future, these errors grow nonlinearly with time.

The planetary and lunar ephemerides DE440 replaces DE430 released in 2014 (Folkner et al. 2014) and its precursors. Since the DE430 release, several interim ephemerides have been released. Each interim DE file was for a specific flight project, which has been tuned for the flight project's target body. For example, the last release was DE438 in 2018 for the Juno mission (Bolton et al. 2017). DE440 has updated all bodies using all available data, including the Moon, which has not been updated since DE430.

Since DE430, several updates have been made to the dynamical model used to integrate DE440. Perturbations from 30 individual Kuiper belt objects (KBOs) and a circular ring representing the rest of the Kuiper belt, modeled as 36 point masses with an equal mass located at 44 au, have been added to the model (Pitjeva & Pitjev 2018). The Lense–Thirring (LT) effect from the Sun's angular momentum has also been added (Park et al. 2017). For the orientation of Earth, the Vondrak precession model has been used (Vondrak et al. 2011), which, according to Vondrak et al. (2009), is more accurate for integrations beyond ±1000 years than the Lieske precession model (Lieske 1979) used for DE430. For the Moon, the effect of geodetic precession on lunar librations has been added as well as the solar radiation pressure force on the Earth–Moon system orbits.

Compared to DE430, new data spanning over about 7 years have been added to compute DE440. The shapes of Mercury, Venus, and Mars orbits are determined mainly by the radio range data of the MErcury Surface, Space ENvironment, GEochemistry, and Ranging (MESSENGER), Venus Express, and Mars-orbiting spacecraft, respectively. The orientations of inner planet orbits are tied to the International Celestial Reference Frame (ICRF) via Very Long Baseline Interferometry (VLBI) of Mars-orbiting spacecraft (Folkner & Border 2015; Park et al. 2015). The orbit of the Moon is determined from laser ranging to lunar retroreflectors. The orbit accuracy of Jupiter has improved substantially by fitting the orbit to the radio range and Very Long Baseline Array (VLBA) data of the Juno spacecraft. The orbit of Saturn is determined by the radio range and VLBA data of the Cassini spacecraft, with improved spacecraft orbits used for processing the radio range data. The orbits of Uranus and Neptune are determined by astrometry and radio range measurements to the Voyager flybys. The orbit of Pluto is now mainly determined by stellar occultations reduced against the Gaia star catalog (Gaia Collaboration et al. 2018; Desmars et al. 2019).

For the Moon, viscous damping between the liquid core and the solid mantle are observed in the lunar laser ranging (LLR) data. This implies an excitation of the relative motion of a lunar core and mantel in the past, possibly due to a spin/orbit resonance that occurred in geologically recent times (Rambaux & Williams 2011). Both DE440 and DE441 have been fit to the same data set, but DE441 assumed no damping between the lunar liquid core and the solid mantle. In this way, a divergence can be avoided when integrated backward for thousands of years. As a consequence, DE441 is less accurate than DE440 for the current century, but more accurate over longer past time spans. DE441 covers years −13,200 to +17,191, compared to DE440 covering years 1550–2650. The difference in the orbits of the planets between DE440 and DE441 are less than 1 m over the 1100 years of the DE440 time span. The difference in the orbit of the Moon between DE440 and DE441 is less than 2 m during the time span of the LLR data, i.e., 1970–2020, but increased over a longer time span, especially in the along-track direction (i.e., velocity direction). Figure 1 shows the difference in the lunar orbit relative to Earth between DE440 and DE441 over 200 years (i.e., DE441 minus DE440). In general, DE440 is recommended for analyzing modern data while DE441 is recommended for analyzing historical data earlier than the modern range data.

Figure 1.

Figure 1. Difference in the lunar orbit relative to Earth between DE440 and DE441 (i.e., DE441 minus DE440) in radial (R), transverse (T), and normal (N) directions.

Standard image High-resolution image

2. Coordinates of Planetary and Lunar Ephemerides

2.1. Inertial Reference Frame

The inertial coordinate frame of the planetary and lunar ephemerides is connected to the International Celestial Reference System (ICRS). The current ICRS realization is achieved by VLBI measurements of the positions of extragalactic radio sources (i.e., quasars) defined in the Third Realization of the International Celestial References Frame (ICRF3; Charlot et al. 2020), which is adopted by the International Astronomical Union (IAU). The orbits of the inner planets are tied to ICRF3 via VLBI measurements of Mars-orbiting spacecraft (Konopliv et al. 2016) with respect to quasars with positions known in the ICRF. Overall, the orientations of inner planet orbits are aligned with ICRF3 with an average accuracy of about 0.2 mas (Folkner & Border 2015; Folkner et al. 2014; Park et al. 2015). The orbits of Jupiter and Saturn are tied to ICRF3 via VLBA measurements of Juno and Cassini spacecraft (Jones et al. 2020), respectively.

2.2. Solar System Barycenter

The solar system barycenter (SSBC) is defined as (Estabrook 1971)

Equation (1)

Here, the summation is over all bodies with finite mass, r i , is the position of body i, and ${\mu }_{i}^{* }$ is defined as

Equation (2)

where GMi is the mass parameter of body i, c is the speed of light, vi is the barycentric speed of body i, and ${r}_{{ij}}=\left|{{\boldsymbol{r}}}_{i}-{{\boldsymbol{r}}}_{j}\right|$ is the distance between bodies i and j.

For DE440, the bodies used for computing the SSBC were the Sun, barycenters of eight planetary systems, the Pluto system barycenter, 343 asteroids, 30 KBOs, and a KBO ring representing the main Kuiper belt. The 343 asteroids were the same set of asteroids used in DE430, which consist of ∼90% of the total asteroid-belt mass. The mass of the 30 largest known KBOs were from Pitjeva & Pitjev (2018). The circular KBO ring was modeled as 36 point masses with equal mass located in the ecliptic plane with a semimajor axis of 44 au, with the ring mass estimated.

Figure 2 shows the motion of the SSBC relative to the Sun for 100 years (2000–2100), which is sometimes called the solar inertial motion (SIM). Compared to DE430, SSBC has shifted by ∼100 km, which is mainly due to the addition of KBOs. It is important to note that, to the first order, Earth orbits around the Sun, not around the SSBC. This point is reflected in Figure 3, which shows the time history of the closest (e.g., perihelion) and farthest (e.g., aphelion) points of the Earth–Moon barycenter (EMB) relative to the Sun and the SSBC. The near-constant distance of the perihelion and aphelion of the EMB relative to the Sun indicates that SIM does not affect the orbit of Earth relative to the Sun.

Figure 2.

Figure 2. Position of the solar system barycenter relative to the Sun in XY (left) and XZ (right) heliocentric ecliptic planes, respectively. The yellow circle represents the Sun.

Standard image High-resolution image
Figure 3.

Figure 3. Distances of the closest and farthest points of the Earth–Moon barycenter relative to the Sun and SSBC of DE440.

Standard image High-resolution image

2.3. Ephemeris Coordinate Time

JPL's DE series are integrated using the barycentric dynamical time (TDB), which is defined relative to the barycentric coordinate time (TCB; Petit & Luzum 2009). All of the data used to compute DE440 and DE441 had the intrinsic time tag in the coordinated universal time (UTC), which differs from the international atomic time (TAI) by leap seconds (i.e., TAI = UTC + leap seconds). In order to process these UTC-tagged measurements, the conversion from UTC to TDB would be needed (Soffel et al. 2003; Petit & Luzum 2009). Once the TAI time is computed, 32.184 s are added to compute the terrestrial time (TT; i.e., TT = TAI + 32.184 s). The conversion from TT to TDB (in Julian days) is given by

Equation (3)

where LG = 6.969290134 × 10−10 defines the rate of TT with respect to geocentric coordinate time (TCG), LB = 1.550519768 × 10−8 defines the rate of TDB with respect to TCB, T0 is 2443144.5003725 Julian days, ${\mathrm{TDB}}_{0}=-65.5\times \tfrac{{10}^{-6}}{86400}$ days, vE is the velocity of the Earth, v E is the velocity vector of the Earth, r E is the position vector of the Earth, and r S is the position vector of a measurement station. The potential term w0E is defined as

Equation (4)

with the summation over all bodies other than Earth. The potential due to external oblate figures wLE is defined as

Equation (5)

where J2⊙ is the unnormalized second-degree gravitational zonal harmonic of the Sun, R is the solar radius, and φE,⊙ is the heliocentric ecliptic latitude of Earth. The term wiE is defined as

Equation (6)

where the summation is over all bodies other than Earth. Lastly, ΔE is defined as

Equation (7)

where a i is the acceleration of body i and the summation is over all bodies.

2.4. Orientation of the Moon

LLR measures the round-trip light time of a laser pulse between an Earth LLR station and a retroreflector on the Moon. Thus, LLR data are not only sensitive to where the Moon is, but they are also sensitive to its orientation.

The orientation of lunar exterior (mantle and crust, hereafter referred to as the mantle) is defined by the principal axes (PAs) of the undistorted lunar mantle, and thus its moment of inertia matrix is diagonal. The directions of the PAs are taken from analyses of the Gravity Recovery and Interior Laboratory (GRAIL) data (Konopliv et al. 2013; Lemoine et al. 2013). The Euler angles that define the rotation from the PA frame to the inertial ICRF3 frame are: ϕm , the angle from the X-axis of the inertial frame along the XY plane to the intersection of the mantle equator; θm , the inclination of the mantle equator from the inertial XY plane; and ψm , the longitude from the intersection of the inertial XY plane with the mantle equator along the mantle equator to the prime meridian. The rotation from the lunar PA frame (i.e., lunar mantle frame) to ICRF3 is given as

Equation (8)

The Euler angles ϕm , θm , and ψm are also known as lunar libration angles, stored in the DE440 and DE441 files. The rotation matrices use the right-hand rule and are defined as

Equation (9)

Equation (10)

Equation (11)

These lunar libration angles are integrated simultaneously with the orbital motion. The equations of motion for the lunar libration angles are

Equation (12)

Equation (13)

Equation (14)

where ωm,x , ωm,y , and ωm,z represent the components of the lunar mantle angular velocity ω m expressed in the mantle frame. The time derivatives of ω m are given in Section 4.

The lunar orientation model includes a fluid core. The orientation of the core with respect to the ICRF is represented by the Euler angles ϕc , θc , and ψc , which are also numerically integrated. Since the shape of the core/mantle boundary is modeled as fixed to the frame of the mantle, it is more convenient to express the core angular velocities with respect to the mantle frame. The time derivatives of the core Euler angles are then given by

Equation (15)

Equation (16)

Equation (17)

where the core angular velocity ω c is related to the angular velocity ${{\boldsymbol{\omega }}}_{c}^{\dagger }$ in a frame defined by the intersection of the core equator with the inertial XY plane by

Equation (18)

The time derivatives of ω c are given in Section 4.

Most of lunar cartographic products are defined relative to the DE421 mean-Earth/mean-rotation (MER) frame. The MER frame is defined by the X-axis pointing toward the mean-Earth direction and the Z-axis pointing toward the mean-rotation axis direction. The rotation from the DE440 PA frame to the DE421 MER frame is estimated by comparing the coordinates of the lunar retroreflectors estimated in the DE440 PA frame and the retroreflector coordinates in the DE421 MER frame, which yield

Equation (19)

Table 1 shows the five lunar retroreflector positions in the DE440 PA frame and the corresponding lunar retroreflector positions in the DE421 MER frame.

Table 1.  XYZ Coordinates of Lunar Retroreflectors in the DE440 PA Frame and the DE421 MER Frame

RetroreflectorsDE440 PA Frame (m)DE421 MER Frame (m)
Apollo 111591967.0491591747.649
 690698.573691222.200
 21004.46120398.110
Apollo 141652689.3691652818.682
 −520998.431−520454.587
 −109729.869−110361.165
Apollo 151554678.1041554937.504
 98094.49898604.886
 765005.863764412.810
Lunokhod 21339363.5981339388.213
 801870.995802310.527
 756359.260755849.393
Lunokhod 11114291.4521114958.865
 −781299.273−780934.127
 1076059.0491075632.692

Download table as:  ASCIITypeset image

2.5. Orientation of Earth

Only the long-term change of the Earth orientation is modeled in the ephemeris integration. The Earth orientation model used for DE440 and DE441 is based on a long-term precession model (Vondrak et al. 2011) and a modified nutation model based on the IAU 1980 precession model including only terms with a period of 18.6 years.

The Earth pole unit vector in the inertial frame, p E, can be computed by the following steps.

First, the mean longitude of the ascending node of the lunar orbit measured on the ecliptic plane from the mean equinox of date is computed by

Equation (20)

where T is the TDB time in Julian centuries (36,525 days) from J2000.0. The nutation angles in longitude, Δψ, and obliquity, Δε, are given by

Equation (21)

Equation (22)

The true pole of date unit vector, p d , is computed by rotating the Earth-fixed pole vector by the effect of the 18.6 year nutation term to give

Equation (23)

where the mean obliquity $\bar{\varepsilon }$ is given by

Equation (24)

The pole unit vector in the inertial frame p E is computed by precessing the pole of date to inertial coordinates using the long-term precession model (Vondrak et al. 2011) plus an estimated frame offset in x and y rotations,

Equation (25)

where Φx and Φy are the estimated offsets of the EME2000 pole from the ICRF pole and the precession matrix ${{ \mathcal R }}_{V}$ is given by

Equation (26)

where k is the ecliptic pole vector and n is the mean equatorial pole vector, derived from polynomial fits to a numerically integrated long-term orientation of Earth (Vondrak et al. 2011). The $\left|\cdot \right|$ operator represents the norm of a vector.

3. Translational Equations of Motion

This section presents the dynamical models of the planetary and lunar ephemerides, including changes and updates made compared to DE430. Some materials from DE430 (Folkner et al. 2014) are repeated so that this paper can be self-contained and the results can be reproduced.

3.1. Point-mass Acceleration

The point-mass interaction between planetary bodies is governed by the parameterized post-Newtonian (PPN) formulation (Will & Nordtvedt 1972; Moyer 2003)

Equation (27)

where the summations are over all bodies, and β and γ are the Eddington–Robertson–Schiff parameters representing the measure of nonlinearity in the superposition law for gravity and the amount of space curvature produced by a unit rest mass, respectively, and are constrained to unity as predicted by the general theory of relativity (GTR).

DE440 integrated the same set of bodies (i.e., Sun, barycenter of eight planets, the Moon, Pluto barycenter, and 343 asteroids) used in DE430, but also included perturbations from 30 KBOs and a KBO ring discussed in Section 2.2. The key mass parameters used in DE440 are shown in Table 2, and all other relevant parameters are given in the comment blocks of the DE440 and DE441 files.

Table 2. Planetary Masses Used in DE440 and DE441

ParameterValue
GMSun 132712440041.279419 km3 s−2 (estimated from DE440)
GMMercury 22031.868551 km3 s−2 (Konopliv et al. 2020)
GMVenus 324858.592000 km3 s−2 (Konopliv et al. 1999)
GMEarth 398600.435507 km3 s−2 (estimated from DE440)
GMMars System 42828.375816 km3 s−2 (Konopliv et al. 2016)
GMJupiter System 126712764.100000 km3 s−2 (SSD JPL 2020)
GMSaturn System 37940584.841800 km3 s−2 (SSD JPL 2020)
GMUranus System 5794556.400000 km3 s−2 (Jacobson 2014)
GMNeptune System 6836527.100580 km3 s−2 (Jacobson 2009)
GMPluto System 975.500000 km3 s−2 (Brozovic et al. 2015)
GMMoon 4902.800118 km3 s−2 (estimated from DE440)
${{GM}}_{\mathrm{Ceres}}$ 62.62890 km3 s−2 (Park et al. 2016; Konopliv et al. 2018; Park et al. 2019, 2020a)
GMVesta 17.288245 km3 s−2 (Konopliv et al. 2014; Park et al. 2014)

Download table as:  ASCIITypeset image

3.2. Point-mass Interaction with Extended Bodies

Nonspherical gravitational interaction has been modeled using a spherical harmonic expansion. For Earth, the interaction of the zonal harmonics up to the fifth degree and the point masses of the Moon, Sun, Mercury, Venus, Mars, Jupiter, and Saturn have been modeled. For the Moon, the interaction of a degree and order 6 gravity field and the point masses of Earth, Sun, Mercury, Venus, Mars, Jupiter, and Saturn have been modeled. For the Sun, the interaction of the second-degree zonal harmonic with all other bodies has been modeled.

The acceleration due to an extended body can be represented as

Equation (28)

where the ξ η ζ coordinate system is defined such that the ξ-axis is defined outward from the extended body to the point mass, the ξ ζ-plane contains the figure spin-pole of the extended body, and the η-axis completes the triad. Here, r is the distance between the two bodies; n1 and n2 represent the maximum degrees of the zonal and nonzonal spherical harmonic coefficients, respectively; Pn and Pnm represent the unnormalized degree-n Legendre polynomial and associated Legendre function with degree-n and order-m, respectively; ${P}_{n}^{{\prime} }$ and ${P}_{{nm}}^{{\prime} }$ represent the derivative of Pn and Pnm with respect to $\sin \varphi $, respectively; Jn represents the degree-n zonal harmonic coefficient; Cnm and Snm represent the nonzonal spherical harmonic coefficients for the extended body; R represents the reference radius of the extended body; and λ and φ represent the longitude and latitude of the point mass in the extended-body fixed coordinate system. Once the accelerations are computed in the body fixed frame, they are transformed into the inertial frame for integration.

There is also an interaction between the figure of an extended body and a point mass, often called the indirect acceleration. Given a i,figi−pmj , which denotes the acceleration of the extended body i interacting with the point-mass external body j expressed in an inertial frame, the corresponding indirect acceleration of the point mass, a j,figi−pmj , is

Equation (29)

For the Moon, the second-degree spherical harmonic coefficients vary with time due to distortion by tides and rotation. These coefficients are computed from the moment of inertia tensors as a function of time and the detailed description is given in Section 4. The second-degree spherical harmonic coefficients are given by

Equation (30)

Equation (31)

Equation (32)

Equation (33)

Equation (34)

where Iij,T represent the elements of the total lunar moment of inertia matrix (see Section 4), mM is the lunar mass, and RM is the lunar radius.

3.3. Acceleration of the Moon from Earth Tides

The lunar orbit is affected by the tides raised on Earth by the Sun and Moon. The tidal distortion of Earth can be modeled using the second-degree gravitational Love numbers, k2j,E , where the order j is 0, 1, and 2 for long-period, diurnal, and semidiurnal responses, respectively. See Section 2.2 in Williams & Boggs (2016) for more information.

A time-delay tidal model has been applied to account for the tidal dissipation. The distorted response of Earth is delayed with respect to the tide-raising forces from the Moon or Sun. The appropriate time delay depends on the period of each tidal component. Consequently, different time delays have been employed for each order j. To allow for time delays shifting across the diurnal and semidiurnal frequency bands, separate time delays are associated with Earth's rotation and the lunar orbit.

The acceleration of the Moon due to the Earth tides is evaluated separately for the tides raised by the Sun and the tides raised by the Moon. The Earth tides depend on the position of the tide-raising body with respect to Earth, r T , where T can denote either the Sun or the Moon. The position of the tide-raising body is evaluated at an earlier time $t-{\tau }_{j}^{{\prime} }$, where ${\tau }_{j}^{{\prime} }$ denotes the orbital time lag, for long-period (j = 0), diurnal (j = 1), and semi-diurnal (j = 2) responses. The rotational distortion of Earth is delayed by a rotational time lag τj , so that the distortion leads the direction to the tide-raising body by an angle $\dot{\theta }{\tau }_{j}$, where $\dot{\theta }$ is the rotation rate of Earth. The long-period zonal tides (j = 0) do not depend on the rotation of Earth, so τj = 0. The acceleration of the Moon due to the distorted Earth depends on the position of the Moon with respect to Earth ( r ) and on the modified position vector for the tide-raising body (${{\boldsymbol{r}}}_{j}^{* }$) that is given for each order j by

Equation (35)

where ${{ \mathcal R }}_{{\rm{E}}}$ rotates the time-delayed position of the tide-raising body with respect to Earth from the inertial frame to the Earth fixed frame and ${{ \mathcal R }}_{z}$ here means a right-handed rotation of the vector ${{\boldsymbol{r}}}_{T}\left(t-{\tau }_{j}^{{\prime} }\right)$ by the angle $-\dot{\theta }{\tau }_{j}$ about Earth's rotation axis.

For evaluation of the acceleration of the Moon, the vectors r (Moon with respect to Earth) and ${{\boldsymbol{r}}}_{j}^{* }$ (time-delayed position of a tide-raising body with respect to Earth) are expressed in cylindrical coordinates with the Z-axis perpendicular to Earth's equator so that r = ρ + z and ${{\boldsymbol{r}}}_{j}^{* }={{\boldsymbol{\rho }}}_{j}^{* }+{{\boldsymbol{z}}}_{j}^{* }$.

The acceleration of the Moon due to the tide raised on Earth by each tide-raising body (Sun or Moon), a M,tide is given by

Equation (36)

where mT is the mass of the tide-raising body.

The tidal acceleration due to tidal dissipation is implicit in the above acceleration. Tides raised on Earth by the Moon do not influence the motion of the EMB. The effect of Sun-raised tides on the barycentric motion is not considered.

The tidal bulge leads the Moon and its gravitational attraction accelerates the Moon forward and retards Earth's spin. Energy and angular momentum are transferred from Earth's rotation to the lunar orbit. Consequently, the Moon moves away from Earth, the lunar orbit period lengthens, and Earth's day becomes longer. Some energy is dissipated in Earth rather than being transferred to the orbit.

3.4. LT Acceleration

In DE440, the acceleration of each body other than the Sun due to the gravito-magnetic effect of GTR, also known as the LT effect, has been implemented (Moyer 2003; Park et al. 2017):

Equation (37)

where the LT angular velocity vector ΩI is given by

Equation (38)

Equation (39)

where G is the universal gravitational constant, γ is the Eddington–Robertson–Schiff parameter from Section 3.1, c is the speed of light, r i is position of the body with respect to the Sun, M is the Sun's mass, C is the Sun's polar moment of inertia divided by $\left({M}_{\odot }{R}_{\odot }^{2}\right)$, where ${C}_{\odot }/({M}_{\odot }{R}_{\odot }^{2})=0.06884$, R is the Sun's equatorial radius (696,000 km), ω the is Sun's rotation rate (14.1844 deg/day), and $\widehat{{\boldsymbol{p}}}$ is the unit spin-pole direction of the Sun (R.A. of 286fdg13 and decl. of 63fdg87; Archinal et al. 2018).

The LT effect is small, but it is important for fitting the MESSENGER range data (Park et al. 2017). Overall, the LT effect causes Mercury's perihelion to precess at the rate of about −0farcs0020/Julian century, which is about 7% of the precession caused by the solar oblateness.

3.5. Solar Radiation Pressure

Photons carry energy and momentum so there is a very small force directed away from the Sun. Like Newtonian gravity, solar radiation pressure depends on the inverse-square distance of the Sun. A simple solar radiation pressure model has been implemented for Earth and the Moon, with accelerations given by

Equation (40)

Equation (41)

where the acceleration due to solar radiation pressure is a very small fraction of gravitational acceleration (Vokrouhlický 1997), i.e., εE,srp = 2 × 10−14 for Earth and εM,srp =1.44 × 10−13 for the Moon. Here, r SE is the inertial Sun-to-Earth position vector and r SM is the inertial Sun-to-Moon position vector.

4. Rotational Dynamics of the Moon

The Moon is modeled as an anelastic mantle with a liquid core. The orientations of the core and mantle are numerically integrated for the core and mantle angular velocities. The angular momentum vectors of the mantle and core are the product of the angular velocities and the moments of inertia. The angular momentum vectors change with time due to torques and due to the distortion of the mantle.

4.1. Rate of Change of Lunar Angular Velocities

In a rotating system, the change in angular velocity ω is related to torques N by

Equation (42)

where I is the moment of inertia tensor. The second term on the right-hand side puts the time derivative into the rotating system. The total lunar moment of inertia I T , which is the sum of the moment of inertia of the mantle I m and the moment of inertia of the core I c , is proportional to the mass mM times the square of the radius RM. Because the fractional uncertainty in the constant of gravitation G is much larger than that for the lunar mass parameter GmM, Equation (42) is evaluated in the integration with both sides multiplied by G.

The components of vectors can be given in the inertial frame, mantle frame, or other frames. Since the moment of inertia matrices are nearly diagonal in the mantle frame, there is great convenience to inverting matrices and performing the matrix multiplications in the mantle frame. The resulting vector components can then be rotated to other frames if desired.

The moment of inertia of the mantle varies with time due to tidal distortions. The distortions are functions of the lunar position and rotational velocities computed at time tτm , where τm is a time lag determined from the fits to the LLR data. The time delay allows for dissipation when flexing the Moon. The time derivative of the angular velocity of the mantle is given by

Equation (43)

where I m is the lunar mantle moment of inertia matrix, N M,figM−pmj is the torque on the lunar mantle due to the point-mass body j, N M,figM−figE is the torque on the lunar mantle due to the interaction between the extended figure of the Moon and the extended figure of Earth, ν gp is the angular rate due to geodetic precession, and N cmb is the torque due to the interaction between lunar mantle and core. The geodetic precession rate is defined as

Equation (44)

where L is the rotation matrix from the mean J2000 frame to the lunar body fixed (i.e., selenographic) frame, v EM is the inertial Earth-to-Moon velocity vector, r EM is the inertial Earth-to-Moon position vector, and r SM is the inertial Sun-to-Moon position vector.

The fluid core is assumed to be rotating like a solid and constrained by the shape of the core–mantle boundary at the interior of the mantle, with the moment of inertia constant in the frame of the mantle. The time derivative of the angular velocity of the core expressed in the mantle frame is given by

Equation (45)

where I c is the lunar core moment of inertia matrix.

4.2. Lunar Moments of Inertia

In the mantle frame, the undistorted moment of inertia of the mantle and the moment of inertia of the core are diagonal. The undistorted total moment of inertia ${\widetilde{{\boldsymbol{I}}}}_{T}$ is given by

Equation (46)

with AT , BT , and CT given by

Equation (47)

Equation (48)

Equation (49)

where mM is the mass of the Moon, RM is the reference radius of the Moon, ${\tilde{J}}_{2,M}$ is the second-degree zonal harmonic of the undistorted Moon, and βL and γL are ratios of the undistorted moments of inertia given by

Equation (50)

Equation (51)

The undistorted total moment of inertia and the second-degree zonal harmonic of the undistorted Moon are not the same as the mean values since the tidal distortions have non-zero averages.

We assume that the lunar rotation aligns an oblate core–mantle boundary with the equator. Then the moment of inertia of the core I c is given by

Equation (52)

where αc = Cc /CT is the ratio of the core polar moment of inertia to the undistorted total polar moment of inertia and fc is the core oblateness. Distortion of the core moment of inertia is not considered.

The undistorted moment of inertia of the mantle is the difference between the undistorted total moment of inertia and the core moment of inertia:

Equation (53)

The moment of inertia of the mantle varies with time due to tidal distortion by Earth and spin distortion,

Equation (54)

where the position vector of Earth relative to Moon, r , and the angular velocity of the mantle, ω m , are evaluated at time tτm ; k2,M is the lunar potential Love number; mE is the mass of Earth; RM is the reference radius of the Moon; r is the Earth–Moon distance; x, y, and z are the components of the position of Earth relative to the Moon referred to the mantle frame; ωm,x , ωm,y , ωm,z are the components of ω m in the mantle frame; and n is the lunar mean motion.

The rate of change of the mantle's moment of inertia is given by

Equation (55)

4.3. Lunar Torques

The torque on the Moon due to an external point-mass A is given by

Equation (56)

where r AM is the position of the point mass relative to the Moon and a M,figM−pmA is the acceleration of the Moon due to the interaction of the extended figure of the Moon with the point-mass A, as described in Section 3.2. Torques are computed for the figure of the Moon interacting with Earth, the Sun, Mercury, Venus, Mars, Jupiter, and Saturn.

It can be shown that torques due to the interaction of the figure of the Moon with the figure of Earth are important for the orientation of the Moon (Eckhardt 1981). The three most significant terms of the torque are

Equation (57)

where ${\widehat{{\boldsymbol{p}}}}_{{\rm{E}}}$ is the direction vector of Earth's pole and ${\widehat{{\boldsymbol{r}}}}_{\mathrm{EM}}$ is the direction vector of Earth from the Moon, I M is the lunar moment of inertia tensor, RE is the reference radius of Earth, and φM,E is defined by $\sin {\varphi }_{{\rm{M}},{\rm{E}}}={\widehat{{\boldsymbol{r}}}}_{\mathrm{EM}}\cdot {\widehat{{\boldsymbol{p}}}}_{{\rm{E}}}$.

The torque on the mantle due to the interaction between the core and mantle is evaluated in the mantle frame and is given by

Equation (58)

where ${\widehat{{\boldsymbol{z}}}}_{m}$ is a unit vector in the mantle frame aligned with the polar axis. Parameter kv is for the viscous interaction at the core–mantle boundary. The torque on the core is the negative of the torque on the mantle.

5. Observational Data Used for Computing DE440 and DE441

The observations that have been used to compute DE440 and DE441 are summarized in Tables 35 for each body.

Table 3. Observational Data for the Moon and Inner Planets

BodyClassificationTypeObservatory/SpacecraftSpanNumber
Moon
 LLRRangeMcDonald 2.7 m1970–19863440
   MLRS/saddle1985–1989275
   MRLS/Mt Fowlkes1988–20142870
   Haleakala1984–1991694
   Observatoire de la Cohat te d'Azur1984–202016425
   Matera2003–2020248
   Apache Point2006–20172452
Mercury
 SpacecraftRangeMariner 101974–19752
   MESSENGER2011–20161353
 Spacecraft3DMESSENGER2008–20103
Venus
 SpacecraftRangeVenus Express2006–20142158
 Spacecraft3DCassini1998–20002
 SpacecraftVLBIMAGELLAN1990–199518
   Venus Express2007–201564
Mars
 SpacecraftRangeViking Lander 11976–19831174
   Viking Lander 21976–197880
   Mars Pathfinder199790
   Mars Express2005–20208751
   Mars Global Surveyor1999–20072130
   Mars Odyssey2002–202010087
   Mars Reconnaissance Orbiter2006–20202634
 SpacecraftVLBIMars Global Surveyor2001–200415
   Mars Odyssey2002–2020169
   Mars Reconnaissance Orbiter2006–2020123
 SpacecraftVLBAVarious2008–20149

Download table as:  ASCIITypeset image

Table 4. Observational Data for Jupiter, Saturn, and Uranus

BodyClassificationTypeObservatory/SpacecraftSpanNumber
Jupiter
 SpacecraftRangeJuno2016–202015
 Spacecraft3DPioneer 1019731
   Pioneer 1119741
   Voyager 119791
   Voyager 219791
   Ulysses19921
   Cassini20001
   New Horizons20071
 SpacecraftVLBAJuno2016–20196
 SpacecraftVLBIGalileo1996–199822
Saturn
 SpacecraftRangeCassini2004–2018147
 SpacecraftVLBACassini2004–201827
 Spacecraft3DVoyager 119801
   Voyager 219811
 AstrometricCCDFlagstaff1998–20163152
   Table Mountain2001–2010687
   Nikolaev1973–1998588
 AstrometricRelativeYerkes1910–192218
Uranus
 Spacecraft3DVoyager 219861
 AstrometricCCDFlagstaff1995–20162362
   Table Mountain1998–2010324
   Nikolaev1961–1999215
   Yunnan2014–20173332
 AstrometricRelativeYerkes1908–192321
 AstrometricTransitBordeaux1985–1993165
   La Palma1984–19971030
   Tokyo1986–198944
   Washington1926–19931783

Download table as:  ASCIITypeset image

Table 5. Observational Data for Neptune and Pluto

BodyClassificationTypeObservatory/SpacecraftSpanNumber
Neptune
 Spacecraft3DVoyager 219891
 AstrometricCCDFlagstaff1995–20152469
   Table Mountain1998–2013416
   Nikolaev1961–1999218
   Yunnan2014–2017755
 AstrometricRelativeYerkes1904–192327
 AstrometricTransitBordeaux1985–1993183
   La Palma1984–19981106
   Washington1926–19931537
Pluto
 AstrometricOccultationVarious1988–201723
 AstrometricCCDFlagstaff1995–20151098
   Table Mountain2001–2015549
   Pico dos Dias1995–20125489
 AstrometricPhotographicPulkovo1930–199253

Download table as:  ASCIITypeset image

An LLR observation measures the round-trip light time from an LLR station on Earth to a retroreflector on the Moon. There are five retroreflectors on the Moon: the Apollo 11, 14, and 15 landing sites and the Lunokhod 2 and 1 rovers. The LLR measurements started in 1970 following the first landings of astronauts and continue to the present. The LLR residuals can be expressed as one-way range residuals, i.e., one-way residual $=\tfrac{\left({t}_{\mathrm{measured}}-{t}_{\mathrm{computed}}\right)c}{2}$. The LLR measurement accuracy has improved with time as technology for producing short-duration high-energy laser pulses and timing measurements has advanced.

Spacecraft measurements are based on the Deep Space Network (DSN) radio range, Doppler, and VLBI measurements. For spacecraft in orbit about the planet, the Doppler measurements are typically used to estimate the position of the spacecraft with respect to the planet's center of mass and then range and VLBI measurements are used to estimate the orbit of the planet. For spacecraft flying by a planet, the range, Doppler, and VLBI data, as available, are used to estimate both the trajectory of the spacecraft and a 3D position of the planet, given as range, R.A., and decl.

Range measurements to spacecraft are usually made at regular intervals during a tracking pass, typically every 10 minutes, while Doppler measurements are made more frequently, typically every minute. Both range and Doppler measurements are based on the measurement of the phase of a radio signal, with the carrier signal used for Doppler and a ranging modulation signal used for range. Since the carrier signal is at a much higher frequency and usually has much higher signal strength, Doppler measurements change in range much more accurately than the range measurements. Because of the shorter wavelength associated with the higher frequency, the integer number of carrier wavelengths cannot be resolved, so Doppler measurements do not allow for an estimation of absolute range. Range measurements are more accurate measurements of round-trip light time. For plotting residuals, one-way residuals are used, which are essentially the same as the LLR residuals, i.e., one-way range $=\,\tfrac{\left({t}_{\mathrm{measured}}-{t}_{\mathrm{computed}}\right)c}{2}$. The range measurement accuracy is often limited by a calibration of the signal path delay in the tracking station prior to each tracking pass. Since this calibration error is common to all range measurements in the tracking pass, there is only one statistically independent range point per pass. Therefore, only one range point per tracking pass was used in the data reduction, and the number of range measurements per spacecraft in Tables 35 reflect this.

Spacecraft VLBI measurements are usually made using two widely separated tracking stations. The measurements are made using a modulation on the carrier signal (delta-differential one-way range) and give one component of the direction to the spacecraft (Thornton & Border 2003). The angular component direction depends on the baseline used. The baseline from Goldstone, California to Madrid, Spain is nearly parallel to Earth's equator, so measurements on that baseline measure an angular component that is close to the R.A.. The baseline from Goldstone, California to Canberra, Australia has an angle of about 45 degrees relative to the equator, thus, it measures an angular component that is approximately mid-way between the R.A. and decl. directions. Residuals for single-baseline measurements are given for each baseline. For Cassini and Juno (and a few points for Mars orbiters), VLBA was used, where the difference in the time of arrival of the spacecraft carrier signal was used to determine both components of the direction to the spacecraft.

Astrometric measurements record the direction to the planet, namely, R.A. and decl., based on imaging relative to a star field. The accuracy of the star catalog is often the largest source of measurement error. The CCD type indicates more modern observations using electronic detectors, generally referred to star catalogs based on the Hipparcos mission launched in 1991 that are referred to the ICRF2 through the estimation of the positions of radio stars using VLBI. Older measurements were taken using photographic plates or transit methods, often referred to older star catalogs, though corrected to the Hipparcos catalog in some fashion. Barnard measured the angular separation between the outer planets and some of their satellites relative to angularly nearby stars at Yerkes Observatory. The positions of those stars are taken from modern star catalogs, with accuracies limited by the knowledge of stellar proper motion. Transit observations cover a longer time span than the more modern spacecraft and astrometric measurements. Since the measurement noise is relatively large for the transit measurements, they do not contribute significantly to the ephemeris solution. The transit measurements are included mainly for historical comparison.

Occultation measurements of Pluto are included here, where the R.A. and decl. are determined from the timed disappearance and reappearance of a star occulted by Pluto (Desmars et al. 2019).

5.1. Residuals

The orbit of the Moon is determined from the LLR data, and Figure 4 shows the one-way residuals of the LLR data over about 50 years. In the last few years, data were available from the Observatoire de la Côte d'Azur, Apache Point, Matera, and Wettzell sites. The rms residual of the early LLR data (∼1970–1980) is about 20 cm while the rms residual of the recent LLR data is about 1.3 cm.

Figure 4.

Figure 4. Residuals of LLR ranges against DE440. The rms residual of the LLR ranges is about 20 cm for the early data and is about 1.3 cm for the recent data.

Standard image High-resolution image

The orbit of Mercury is mainly determined from the MESSENGER radio range data, and Figure 5 shows the one-way range residuals over about 4 years. The rms residual of the MESSENGER ranges is about 0.7 m.

Figure 5.

Figure 5. Residuals of the MESSENGER range data against DE440. The rms residual of the MESSENGER ranges is about 0.7 m.

Standard image High-resolution image

The orbit of Venus is mainly determined from the Venus Express radio range data, and Figure 6 shows the one-way range residuals over about 6.5 years. The rms residual of the Venus Express ranges is about 8 m.

Figure 6.

Figure 6. Residuals of the Venus Express range data against DE440. The rms residual of the Venus Express ranges is about 8 m.

Standard image High-resolution image

The shape of Mars' orbit is mainly constrained by the radio range data of Mars orbiters, and Figure 7 shows the one-way range residuals of Mars Express (MEX), Mars Global Surveyor (MGS), Mars Odyssey (ODY), and Mars Reconnaissance Orbiter (MRO). The rms residual of the MEX ranges is about 2 m and the rms residuals of the MGS, ODY, and MRO ranges are about 0.7 m. The orientation of Mars' orbit is mainly constrained by VLBI measurements of the Mars orbiters, and Figure 8 shows the VLBI residuals of the Goldstone–Madrid and Goldstone–Canberra baselines. The rms residual of the Goldstone–Madrid baseline is about 0.25 mas and the rms residual of the Goldstone–Canberra baseline is about 0.18 mas.

Figure 7.

Figure 7. Residuals of the Mars orbiter range data against DE440. The rms residual of the MEX ranges is about 2 m and the rms residuals of the MGS, ODY, and MRO ranges are about 0.7 m.

Standard image High-resolution image
Figure 8.

Figure 8. Residuals of Mars orbiter VLBI data against DE440, showing the Goldstone–Madrid baseline (top) and the Goldstone–Canberra baseline (bottom). The rms residual of the Goldstone–Madrid baseline is about 0.25 mas and the rms residual of the Goldstone–Canberra baseline is about 0.18 mas.

Standard image High-resolution image

The shape of Jupiter's orbit is mainly constrained by the Juno radio range data, and Figure 9 shows the one-way range residuals over about 4 years. The rms residual of the Juno ranges is about 13 m. The orientation of Jupiter's orbit is mainly constrained by the VLBA data of the Juno spacecraft, and Figure 10 shows the Juno VLBA residuals over about 2 years. The rms residual of the Juno VLBA data in R.A. is about 0.4 mas and the rms residual of the Juno VLBA data in decl. is about 0.6 mas.

Figure 9.

Figure 9. Residuals of the Juno range data against DE440. The rms residual of the Juno ranges is about 13 m.

Standard image High-resolution image
Figure 10.

Figure 10. Residuals of the Juno VLBA data against DE440. The rms residual of the Juno VLBA data in R.A. is about 0.4 mas and the rms residual of the Juno VLBA data in decl. is about 0.6 mas.

Standard image High-resolution image

The shape of Saturn's orbit is mainly constrained by the Cassini radio range data, and Figure 11 shows the one-way range residuals over about 13 years. The rms residual of the Cassini ranges is about 3 m. The orientation of Saturn's orbit is mainly constrained by the VLBA data of the Cassini spacecraft, and Figure 12 shows the Cassini VLBA residuals over about 13 years. The rms residual of the Cassini VLBA data in R.A. is about 0.6 mas and the rms residual of the Cassini VLBA data in decl. is about 0.8 mas if the two obvious outliers are included, and 0.35 and 0.36 mas if they are excluded.

Figure 11.

Figure 11. Residuals of the Cassini range data against DE440. The rms residual of the Cassini ranges is about 3 m.

Standard image High-resolution image
Figure 12.

Figure 12. Residuals of the Cassini VLBA data against DE440. The rms residual of the Cassini VLBA data in R.A. is about 0.6 mas and the rms residual of the Cassini VLBA data in decl. is about 0.8 mas if the two obvious outliers are included, and 0.35 and 0.36 mas if they are excluded.

Standard image High-resolution image

The orbits of Uranus and Neptune are mainly constrained by astrometry and radio range measurements to the Voyager flybys and they are statistically consistent with DE430 (see Folkner et al. 2014).

The orbit of Pluto is mainly determined by astrometry, and Figure 13 shows the residuals of the stellar occultation measurements from Desmars et al. (2019). The rms residuals of the stellar occultations are about 8 mas in R.A. and about 11 mas in decl.

Figure 13.

Figure 13. Residuals of stellar occultations of Pluto against DE440. The rms residuals of the stellar occultations are about 8 mas in R.A. and about 11 mas in decl.

Standard image High-resolution image

6. Conclusions

This paper presents JPL's new general-purpose ephemerides DE440 and DE441 created by fitting numerically integrated orbits to ground-based and space-based observations. Compared to DE430, the previous general-purpose ephemerides released in 2014, new data spanning over about 7 years have been added to compute DE440. The shapes of the orbits of Mercury, Venus, and Mars are constrained mainly by the radio range data of the MESSENGER, Venus Express, and Mars-orbiting spacecraft, respectively. The orientations of inner planet orbits are tied through the ICRF via VLBI of Mars-orbiting spacecraft. The orbit of the Moon is primarily determined by laser ranging to lunar retroreflectors with the data arc extended through 2020 March. The orbit accuracy of Jupiter has improved substantially by fitting the orbit to the new Juno radio ranges and VLBA data. The orbit of Saturn is determined by radio ranges and VLBA data of the Cassini spacecraft, with improved calibration of the radio range data. The orbits of Uranus and Neptune are constrained by astrometry and radio range measurements to the Voyager flybys. The orbit of Pluto is constrained by stellar occultation data reduced against the Gaia star catalog. DE440 spans years 1550–2650 and DE441 spans years −13,200 to +17,191. The ephemerides of DE440 are recommended for analyzing modern data while DE441 is recommended for analyzing historic data before the DE440 time span.

The authors would like to thank A. Konopliv, R. Jacobson, J. Border, D. Jones, T. Morely, and F Budnik for providing some of the data used to compute DE440 and DE441. This research was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. Government sponsorship is acknowledged.

Please wait… references are loading.
10.3847/1538-3881/abd414