26
THE DETERMINATION OF GRAVITATIONAL POTENTIAL DIFFERENCES FROM SATELLITE-TO-SATELLITE TRACKING Christopher Jekeli Department of Civil and Environmental Engineering and Geodetic Science The Ohio State University 2070 Neil Ave. Columbus, OH 43210 e-mail: [email protected] revision submitted to Celestial Mechanics and Dynamical Astronomy 11 October 1999

THE DETERMINATION OF GRAVITATIONAL POTENTIAL ......Keywords: gravitational potential, satellite-to-satellite tracking, range-rate measurements, Earth rotation. 1. INTRODUCTION A satellite

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • THE DETERMINATION OF GRAVITATIONAL POTENTIAL

    DIFFERENCES FROM SATELLITE-TO-SATELLITE TRACKING

    Christopher Jekeli

    Department of Civil and Environmental Engineering and

    Geodetic ScienceThe Ohio State University

    2070 Neil Ave.Columbus, OH 43210

    e-mail: [email protected]

    revision submitted to

    Celestial Mechanics and Dynamical Astronomy

    11 October 1999

  • ABSTRACT

    A new, rigorous model is developed for the difference of gravitational potential between two close

    Earth-orbiting satellites in terms of measured range-rates, velocities and velocity differences, and

    specific forces. It is particularly suited to regional geopotential determination from a satellite-to-

    satellite tracking mission. Based on energy considerations, the model specifically accounts for the

    time variability of the potential in inertial space, principally due to Earth’s rotation. Analysis

    shows the latter to be a significant ( ± 1 m2/s2 ) effect that overshadows by many orders of

    magnitude other time dependencies caused by solar and lunar tidal potentials. Also, variations in

    Earth rotation with respect to terrestrial and celestial coordinate frames are inconsequential. Results

    of simulations contrast the new model to the simplified linear model (relating potential difference to

    range-rate) and delineate accuracy requirements in velocity vector measurements needed to

    supplement the range-rate measurements. The numerical analysis is oriented toward the scheduled

    Gravity Recovery And Climate Experiment (GRACE) mission and shows that an accuracy in the

    velocity difference vector of 2×10–5 m/s would be commensurate within the model to the

    anticipated accuracy of 10–6 m/s in range-rate.

    Keywords: gravitational potential, satellite-to-satellite tracking, range-rate measurements, Earth

    rotation.

    1. INTRODUCTION

    A satellite mission dedicated to the improvement of our knowledge of the Earth’s gravitational field

    with a direct (in situ) measurement system has been in the proposal stages for a long time and at

    - 1 -

  • several agencies. Of course, gravitational field knowledge comes also by tracking satellites from

    ground stations, and many long-wavelength models of the field have been deduced from such data.

    But, these models derive from the observations of a large collection of satellites that have been

    tracked over various periods during the long history of Earth-orbiting satellites, where none of

    these was launched for the expressed purpose of providing a global and detailed model of the

    gravitational field.

    Rather, the proposed gravity mapping missions are based on one of several related

    measurement concepts, including the measurement of the range between two close Earth-orbiting

    satellites (GRAVSAT, GRM: Keating et al., 1986), tracking a low-orbiting satellite with a system

    of high-orbiting satellites (Jekeli and Upadhyay, 1990), or measuring the gravitational gradients on

    a single low-orbiting satellite (ARISTOTELES: Bernard and Touboul, 1989; SGGM: Morgan and

    Paik, 1988; GOCE, Gravity Field and Steady-State Ocean Circulation Explorer: Rummel and

    Sneeuw, 1997).

    Such a mission now has been approved and is expected to be realized in 2001. GRACE, the

    Gravity Recovery And Climate Experiment (Tapley and Reigber, 1998) is a variant of the erstwhile

    GRAVSAT and GRM mission concepts in that two low-altitude satellites will track each other as

    they circle the Earth in identical near polar orbits. Unlike GRM, the satellites are not “drag-free”

    and non-gravitational accelerations must be measured independently using on-board

    accelerometers. Also, the altitude of the GRACE satellites is significantly higher (400 km) than

    that proposed for GRM (160 km). Another significant departure from the previous concept is that

    each satellite will carry a geodetic quality GPS receiver. The purpose of these receivers is to aid in

    orbit determination, as well as provide GPS satellite occultation measurements to model the lower

    atmosphere.

    As shown also here with equation (23), a simple model may be derived on the basis of energy

    conservation that relates the measured range-rate between two satellites to the gravitational potential

    difference. However, though widely used to analyze the capability of a satellite-to-satellite tracking

    - 2 -

  • (SST) mission to determine the geopotential (Wolff, 1969; Jekeli and Rapp, 1980; Wagner, 1983;

    Dickey, 1997), it is hardly adequate as a model for processing actual data. In fact, this model

    neglects the significant effect of Earth’s rotation that causes the geopotential to vary with time in

    inertial space (therefore, strictly, it is non-conservative). Furthermore, the range-rate accounts for

    but a single component of the velocity vector difference resulting from the potential difference.

    These deficiencies in the model are orders of magnitude above the measurement noise level and

    would preclude accurate in situ geopotential determination. It should be noted, however, that other

    modeling techniques exist to determine the geopotential on global and regional bases. For

    example, the range-rate or range may be expressed in terms of a spherical harmonic series of the

    global geopotential (Colombo, 1984) or locally in terms of suitable basis functions (Ilk, 1986), and

    the corresponding coefficients are solved using a least-squares adjustment procedure.

    The in situ model developed here is particularly suited to regional determination of the

    geopotential and would also be amenable to global determination using conventional harmonic

    analysis techniques. It is based on an energy equation generalized to account for the time-varying

    potential fields. Results of simulations show the relationship between geopotential accuracy and

    accuracies in range-rate and velocity vector measurements associated with the GRACE mission.

    Clearly, this model applies to any other SST mission to map the gravitational field of any planet.

    2. THE MODEL

    From energy considerations (see the Appendix), the exact relationship in inertial space between the

    gravitational potential, V, and terms containing the satellite velocity, x = x1, x2, x3 , and specific

    forces acting on the satellite, F = F1, F2, F3 , is given by (A.14) with (A.5) substituted:

    - 3 -

  • V = 12 x

    2– Fk xk dt

    t0

    t

    Σk

    + ∂V∂t dtt0

    t

    – E0 (1)

    The first term on the right hand side is the kinetic energy and the second term represents energy

    dissipation. The third term is due to the explicit time variation of the gravitational potential in

    inertial space; and E0 is the energy constant of the system.

    If we measure a satellite’s velocity along its orbit, as well as the action forces on the satellite,

    then (1) represents an (integral) equation that can be solved for the potential, V. We decompose

    the potential as follows:

    V = Vrotating Earth + Vlunar tide + Vsolar tide + Vplanetary tides

    + Vsolid Earth tide + Vocean tide + Vatmospheric tide

    + Vocean loading + Vatmospheric loading + Vother mass redistributions

    (2)

    and recognize that some parts are better known than others and most have dissimilar magnitudes

    and periodicities. The gravitational potential of the rotating Earth can be expressed in spherical

    polar coordinates in an Earth-fixed coordinate frame using spherical harmonic functions, Yn,m :

    Vrotating Earth ≡ Ve(r,θ,λ) =

    kMeR Σn = 0

    ∞ Rr

    n + 1Cn,m Yn,m(θ,λ)Σm = – n

    n(3)

    where r is geocentric radius, θ is co-latitude, and λ is longitude with respect to a defined zero-

    meridian; kMe is the gravitational constant times Earth’s total mass (including atmosphere); R is a

    mean Earth radius; Cn,m are coefficients that define Earth’s mass density distribution; and

    - 4 -

  • Yn,m(θ,λ) = Pn, m (cosθ)cosmλ, m ≥ 0

    sin m λ, m < 0(4)

    where Pn,m are fully normalized associated Legendre functions. The frame for the coordinates

    (r,θ,λ) is fixed to the Earth and it realizes the International Terrestrial Reference System that is

    well defined by the International Earth Rotation Service (IERS) (McCarthy, 1996). The

    coefficients, Cn,m , are assumed constant since any temporal redistribution of mass is accounted

    for by the other potential components in (2).

    The potential in (1) is supposed to be in the inertial frame. Hence, using (3) requires a

    transformation from the fixed terrestrial to the inertial (mean celestial) frame. It is convenient to

    describe this transformation in terms of co-latitude and longitude angles:

    θ = ζ + ∆ζP + ∆ζN + ∆θS (5)

    λ = α + ∆αP + ∆αN – ωet + ∆λS (6)

    where the coordinates (ζ,α) are the co-declination and right ascension in the inertial frame of

    epoch J2000.0. The terms ∆θS and ∆λS rotate the terrestrial pole of date to the celestial pole of

    date using coordinates of polar motion; ωe is Earth’s rate of rotation and the corresponding term in

    (6) rotates the terrestrial frame into the celestial about the 3-axis by the Greenwich sidereal time;

    ∆ζN and ∆αN account for the nutations of the celestial pole and transform it from its true to its

    mean direction of date; and ∆ζP and ∆αP describe the precession of the pole from its mean

    direction of date to its mean direction at a defined epoch, currently J2000.0. Detailed expressions

    for these terms can be found in (Mueller, 1969) and (Seidelmann, 1992). Each one is an explicit

    function of time, meaning that if (5) and (6) are substituted into (3), then Ve as a function of (ζ,α)

    - 5 -

  • depends explicitly on time.

    These transformations sometimes are interpreted to cause time dependencies in the harmonic

    coefficients; however, in the sequel the present interpretation of time-dependent coordinates is

    preferred and more appropriate. In this way the explicit time-derivative of the potential is given by

    ∂Ve∂t =

    ∂Ve∂θ

    ∂θ∂t +

    ∂Ve∂λ

    ∂λ∂t (7)

    where ∂θ ∂t∂θ ∂t and ∂λ ∂t∂λ ∂t denote explicit time derivatives of these coordinates (now in the inertial

    frame). We note that the dominant explicit time-derivative component in (5) and (6) is – ωe . In

    fact, the precession rates in right ascension and in declination are less than 50 arcsec per year, or

    less than 8×10–12 rad/s . Similarly the nutation rates in longitude and ecliptic obliquity are less

    than 3×10–12 rad/s , and polar motion rates are less than 3×10–13 rad/s for the main Chandler

    wobble. These rates are seven to eight orders or magnitude smaller than

    ωe = 7.292115×10–5 rad/s , and we may approximate

    ∂Ve∂t = – ωe

    ∂Ve∂α (8)

    where, because of the linear relationship (6), ∂ ∂λ∂ ∂λ = ∂ ∂α∂ ∂α . Furthermore, we assume to a similar

    level of approximation that ωe is constant.

    The other potential terms in (2) may be analyzed similarly. Expressed in inertial frame

    coordinates with origin at Earth’s center of mass, the tidal potential of an extra-terrestrial body,

    including the indirect effect arising from the consequent deformation of the quasi-elastic Earth, is

    given approximately by (Torge, 1991; Lambeck, 1988)

    - 6 -

  • VB(r,θ,α) =

    34

    kMBrB

    rrB

    21 + k 2

    Rr

    5⋅

    sin2θ sin2θB cos(α – αB) + sin2θ sin2θB cos2(α – αB) + 3 cos2θ –13 cos

    2θB –13

    (9)

    where kMB is the gravitational constant times the mass of the body, (rB,θB,αB) are its

    coordinates in the inertial frame, and k 2 = 0.29 is Love’s number (an empirical number based on

    observation). Equation (9) treats the body as a point mass and neglects terms with powers in r rBr rB

    greater than 2, which is adequate in the present context for the most influential bodies, the sun and

    the moon. Also, it is assumed that the elastic response to the tidal potential is instantaneous. In

    reality there is a lag, which to a first approximation is constant and, therefore, presently of no

    consequence.

    The coordinates (rB,θB,αB) are all functions explicitly of time due to the motion of the body

    with respect to the Earth. However, the largest rate is in αB since the sun and moon, respectively,

    depart by at most 23.°5 and 29° in declination from the equatorial plane. If we ignore the time

    dependence of rB , then

    ∂VB∂t ≈ –

    ∂VB∂α αB +

    ∂VB∂θB

    θB (10)

    again, because ∂ ∂αB∂ ∂αB = – ∂ ∂α∂ ∂α . If nB denotes the mean angular motion of the body, then the rate

    in co-declination varies between zero and ± sin(i) nB , the latter occurring when the body crosses

    the celestial equator, where i is the inclination of its orbit. The length of a sidereal month is

    approximately 27 days, hence, for the moon, nM = 2.7×10–6 rad/s . The sidereal year is about 365

    days long, implying that the sun’s mean motion is nS = 2.0×10–7 rad/s . The corresponding rates

    in right ascension, αM and αS , have the same respective orders of magnitude. These rates are 1

    - 7 -

  • to 2 orders of magnitude less than Earth’s rate of rotation.

    Evaluating the first term (Doodson’s constant) in (9) for the sun and the moon, we find with a

    satellite altitude of 400 km:

    34

    kMBrB

    rrB

    2= 3.0 m

    2/s2 , moon1.4 m2/s2 , sun

    (11)

    These potentials are smaller than Earth’s gravitational potential by seven orders of magnitude.

    Since the corresponding gradients compare similarly, we have

    O

    ∂VB∂t < 10

    –8 O∂Ve∂t (12)

    for the principal bodies, moon and sun; the effects of other planets may be ignored.

    Lambeck (1988) also gives the potential due to the ocean tides (including the loading effect on

    the solid Earth) and states that the amplitudes are less than 15% of the solid Earth tidal effect that is

    included in (9). On the basis of these magnitudes we may safely neglect these as well as all other

    potentials in (2) as far as the explicit time derivative is concerned; and we have from (8):

    ∂V∂t ≈ – ωe

    ∂Ve∂α (13)

    Now, since x1 = r cosθ cosα and x2 = r cosθ sinα , it is readily shown that

    ∂V∂t = – ωe x1

    ∂Ve∂x2

    – x2∂Ve∂x1

    (14)

    Substituting (A.12) we then have

    - 8 -

  • ∂V∂t = ωe x1 F2 +

    ∂δV∂x2

    – x2 – x2 F1 +∂δV∂x1

    – x1 (15)

    where, from (2), V = Ve + δV . Again, the gradients of the perturbing potential, δV , are about

    seven orders of magnitude less than the acceleration of the satellite, and in most cases so are

    accelerations associated with the atmospheric drag and solar radiation pressure that constitute F .

    Neglecting these terms, we have

    ∂V∂t ≈ ωe x2 x1 – x1 x2 (16)

    which yields

    ∂V∂t dt

    t0

    t

    = – ωe x1 x2 – x2 x1 (17)

    (the constant of integration is relegated to E0 ). As an aside, (17) can also be written as

    ∂V∂t dt

    t0

    t

    = – ωe α x12 + x2

    2 (18)

    This differs from the usual “rotation potential” found in textbooks on celestial mechanics. The

    difference is that here the potential is given in the inertial frame, whereas the rotation potential,

    ωe2 x12 + x2

    2 (see, e.g., Danby, 1988), applies to the Earth-fixed (rotating) frame. To distinguish

    our term, we call it the “potential rotation” term, since it accounts for the rotation of the potential in

    the inertial frame.

    Finally, we arrive at the model for the potential from (1) and (17):

    - 9 -

  • V = 12 x

    2 – Fk xk dtt0

    t

    Σk

    – ωe x1 x2 – x2 x1 – E0 (19)

    This expresses the desired gravitational potential in terms of measured quantities, specific force and

    velocity (also satellite position is required, but not to extremely high accuracy for the potential

    rotation term). The model is approximate only because certain time dependencies in the

    gravitational potential have been neglected according to (16). The energy dissipation is not

    negligible, being of approximately the same order as the potential rotation term. However, it is

    ignored at present to simplify the subsequent analysis.

    3. SATELLITE-TO-SATELLITE TRACKING

    Satellite-to-satellite tracking, for example, as proposed for the GRACE mission, constitutes the

    very precise measurement of the range, ρ12 , between two satellites following each in

    approximately the same orbit. We have ρ12 = e12T

    x12 , where x12 = x2 – x1 , and e12 is the unit

    vector identifying the direction to the second satellite from the first. Then, the range-rate, being

    derived from the measured range, is the projection of the velocity difference between the satellites

    onto the line joining them:

    ρ12 = e12T

    x12 (20)

    since e12T

    e12 = 0 . We treat the range-rate as the measurement, noting that it is only a component

    of the velocity difference.

    For satellites in drag-free orbits ( F = 0 ) and a static gravitational field ( ωe = 0 ), the energy

    - 10 -

  • equation (19) reduces to

    V = 12 x2 – E0 (21)

    Taking the along-track derivative, denoted by da , on both sides yields

    daV = xT

    dax (22)

    If the two satellite are close then the left side may be interpreted as the difference in gravitational

    potential between the satellites and the along-track differential velocity as the range-rate, thus:

    V2 – V1 ≡ V12 ≈ x1 ρ12 (23)

    This relates the measurements directly to potential differences along the orbit. It is the model

    assumed in the analyses by Wolff (1969), Fischell and Pisacane (1978), Rummel (1980), Jekeli

    and Rapp (1980), Wagner (1983), and Dickey (1997), among others.

    Up to the approximations discussed in connection with (19), the correct expression is given by

    V12 = x1

    Tx12 +

    12 x12

    2 – F2k x2k – F1k x1k dtt0

    t

    Σk

    – ωe x121 x22 – x22 x121 – x11 x122 + x122 x11 – E012

    (24)

    where the first two terms derive from x22 – x1

    2 = x2 – x1T

    x2 + x1 , and E012 is a

    constant. Omitting the dissipative term, we write

    V12 = x1T

    x12 +12 x12

    2 + VR12 – E012 (25)

    - 11 -

  • with VR12 denoting the difference in potential rotation terms.

    It is customary to introduce a known reference potential that accounts for the longest

    wavelengths of the signal. We denote all quantities referring to such a reference field by the

    superscript “0”; and by definition, it and all associated quantities, in particular the corresponding

    orbital reference ephemerides of both satellites, can be computed without error. The reference field

    may be a potential with just the central and second zonal harmonic terms; or it may be a low-degree

    spherical harmonic expansion of the potential, say, complete to degree and order 10. For the

    present purposes, a harmonic expansion complete to degree and order 2 will suffice to provide a

    reasonably quantitative illustration. The residual to any of the reference quantities is denoted with

    the prefix “ ∆ ”.

    It must be emphasized that a residual quantity is the difference between a quantity that refers to

    the actual orbit and a quantity that refers to a reference orbit. That is, the only common coordinate

    between the two is time, and not position. Figure 1 illustrates this situation. It is assumed that

    there is a point in time when the two orbits are tangent (i.e., their Keplerian elements coincide).

    Reference orbitTrue Orbit

    V2

    V1

    V02

    •••

    V01

    Figure 1: The geometry of residual quantities referred to a reference orbit.

    The residual quantities are, for example, ∆V12 = V12 – V120

    , ∆x1 = x1 – x10

    , and

    ∆ρ12 = ρ12 – ρ120

    , where the reference potential (sans dissipative energy term) is given analogous

    - 12 -

  • to (25) by:

    V12

    0 = x10 T

    x120

    + 12 x120 2

    + VR120 – E0

    0

    12(26)

    and the residual potential difference is

    ∆V12 = x1T

    x12 – x10 T

    x120

    + x120 T ∆x12 +

    12

    ∆x12T ∆x12 + ∆VR12 – ∆E012 (27)

    Corresponding to the approximation (23), we define the approximate residual model,

    designated with the symbol “^” as

    ∆V12 = x10 ∆ρ12 (28)

    The error in this model relative to the true model (27) is given by

    ε∆V12 = ∆V12 – ∆V12

    = x20

    – x10

    e12T

    ∆x12 + ∆x1 – x10 ∆e12

    T

    x120

    + x1T ∆x12 +

    12

    ∆x122

    – ∆VR12 + ∆E012

    = ν1 + ν2 + ν3 + ν4 – ∆VR12 + ∆E012

    (29)

    which is readily derived using ∆ρ12 = ρ12 – ρ120

    = e12T

    x12 – e120 T

    x120

    . Equation (28) also

    provides an approximate relationship between the error in potential difference resulting from an

    error in the satellite-to-satellite range-rate measurement. Since the velocity magnitude is

    approximately x10

    = 7700 m/s , a standard deviation in the range-rate measurement of 10–6 m/s

    - 13 -

  • (to be expected for the GRACE mission) is equivalent to a standard deviation of about

    0.008 m2/s2 in the potential difference.

    4. A SIMULATION

    To quantify the terms in the error of the potential difference model (28), the orbits of two satellites

    were generated on the basis of the high-degree ( nmax = 360 ) spherical harmonic model of the

    geopotential, EGM96 (Lemoine et al., 1998), but only up to degree and order 180:

    V(r,θ,λ) = kM

    R Σn = 0180 R

    rn + 1

    Cnm Ynm(θ,λ)Σm = –nn

    (30)

    This model was substituted into (A.6) (with F = 0 ) and equation (A.8) was integrated by the

    Adams-Cowell multistep predictor-corrector algorithm yielding the ephemeris (x and x ) of each

    satellite at one-second intervals. The accuracy of the numerical integration of (A.8) was checked

    by comparing the potential difference obtained from (25) to the original difference on the basis of

    (30) — the disagreement over a single revolution was near the limit of the computational precision.

    Other parameters of the two orbits include an initial altitude of 400 km above the Earth’s mean

    radius, an initial eccentricity of zero, and an initial inclination to the equator of 87°; hence they are

    near-polar orbits. The initial orbital elements of the two satellites were chosen so that their

    separation was about 200 km and the two orbital paths never deviated from each other by more

    than 60 m, mostly in the radial direction. The orbital integration was limited to slightly more than a

    single revolution of the satellite pair (about 6000 s). Also, a pair of reference orbits was generated

    using a potential field complete to degree and order 2. The resulting residual potential difference

    between the two satellites was on the order of ± 30 m2/s2 .

    This signal and the error in the model (28) are both shown in Figure 2 for the special case of

    - 14 -

  • identical orbits for the two satellites, meaning that the gravitational potential was assumed to be

    static ( ωe = 0 , for this case, only). In this case, the terms, ν1 and ν2 , on the right side of (29)

    nearly cancel and the model error is three orders of magnitude smaller than the signal. However,

    when the orbits are only similar (within 60 m, and ωe ≠ 0 ), the model error is as large as the signal

    itself (Figure 3), but has a very long-wavelength (once-per-revolution) structure that is caused by

    the second term, ν2 , in (29), as seen in Figure 4. Thus, the stratagem of using the along-track

    derivative to develop the model is rather sensitive to the radial similarity of the orbits.

    ∆V12

    time [s]

    m2 /

    s2

    ∆V12 – ∆V12 × 103

    Figure 2: Comparison of true residual potential difference to model (41) (no Earth

    rotation, identical orbits)

    - 15 -

  • time [s]

    m2 /

    s2

    ∆V12∆V12

    ^

    Figure 3: Comparison of true residual potential difference to model (41) (Earth rotation,

    unequal orbits differing by less than 60 m)

    Figure 5 shows the other model errors associated with the simple model (28). Term ν3 has

    the same order of magnitude as the error due to range-rate measurement error ( 10–6 m/s ), and ν4

    is practically negligible; but the potential rotation term, ∆VR12 , on the order of ± 1 m2/s2 , is

    significant. Therefore, the accuracy of the model (28) is not consistent with a measurement

    accuracy of 10–6 m/s . This means that range-rates cannot be used to full advantage to measure

    potential differences, unless supplemented by velocity vector measurements.

    - 16 -

  • ν2

    ν1 x 10

    time [s]

    m2 /

    s2

    Figure 4: Model error terms ν1 and ν2 for the case depicted in Figure 3.

    ν4 x 104

    ν3 x 103

    Potential Rotation Term

    time [s]

    m2 /

    s2

    Figure 5: Model error terms ν3 , ν4 , and ∆VR12 for the case depicted in Figure 3.

    The more accurate model for the determination of potential differences (again, omitting the

    dissipative energy term), given that range-rates are the primary measurements, is obtained from

    (29) and (28) as:

    ∆V12 = x10 ∆ρ12 – ν1 – ν2 – ν3 – ν4 + ∆VR12 – ∆E012 (31)

    It requires also measurements of velocity vectors and their intersatellite differences. The constant,

    - 17 -

  • ∆E012 , is either obtained from known initial conditions or determined empirically as a bias from a

    sufficiently long sequence of data. To measure a satellite’s velocity generally requires extensive

    ground tracking to determine its ephemeris. However, if the two satellites are equipped with

    Global Positioning System (GPS) receivers (as in the case of the GRACE satellites), then their

    relative velocities can be measured in situ using standard baseline determination procedures

    developed for terrestrial kinematic applications where the current accuracy is estimated to be about

    1 cm/s. In space, the accuracy would be significantly better since the signals transmitted from the

    GPS satellites are unaffected by tropospheric delays. Also, if the clock errors of the GPS satellites

    are known, then the absolute velocity of either satellite can be determined quite accurately (in fact,

    GPS will be used for precise orbit determination of GRACE).

    Nevertheless, the accuracy requirements are rather demanding when measuring velocities and

    velocity differences associated with the potential difference determination according to (31).

    Figure 6 shows the relationship between the accuracy in potential difference, δ∆V12 , and

    accuracies in range-rate ( δρ12 ), absolute ( δx1 ), and intersatellite ( δx12 ) velocity measurements.

    The principal term affected by errors in absolute velocity is ν2 ; while the velocity difference error

    affects mostly the potential rotation term.

    Computation of these two terms also requires accurate absolute (for ∆VR12 ) and relative (for

    ν2 ) position vector measurements. Figure 7 shows the corresponding relationships to the

    potential difference accuracies. For example, determination of the potential difference along the

    satellite trajectory to an accuracy of 0.1 m2/s2 (corresponding to an accuracy of 1 cm in geoid

    differences) requires accuracies in range-rate, velocity, and position as follows:

    δρ12 = 1×10–5 m/s , δx1 = 5×10

    –4 m/s , δx12 = 2×10–4 m/s

    δx1 = 7 m , δx12 = 1×10–2 m(32)

    The vector position requirements are easily satisfied with GPS, while the velocity vector

    - 18 -

  • requirements are just beyond current demonstrated GPS capability, but not outside the realm of

    feasibility. Note that the anticipated order-of-magnitude higher accuracy in range-rate for GRACE

    would be advantageous only with commensurate improvements in velocity and position accuracies.

    δ∆V12 [m2/s2]

    [m/s

    ]

    δx1 δx12

    δρ12

    Figure 6: Range-rate and velocity accuracy requirements for potential difference

    determination according to (31).

    - 19 -

  • δ∆V12 [m2/s2]

    [m]

    δx1

    δx12

    Figure 7: Position accuracy requirements for potential difference determination according

    to (31).

    5. SUMMARY

    An accurate model for the gravitational potential difference was developed for the satellite-to-

    satellite tracking system concept. The model relates potential difference to in situ measurements of

    velocity (consisting of range-rate, relative and absolute velocity vectors), position, and specific

    force. In particular, the model includes the time dependencies of the gravitational potential in

    inertial space, dominated for practical purposes by Earth’s constant rotation rate. Moreover, the

    model also differs from models usually used by terms that depend on the velocity difference

    vector. Simulations show that the accuracy of this velocity difference is allowed to be about one

    order of magnitude poorer than the range-rate accuracy. They also show that the potential rotation

    term is significant at the level of 1 m2/s2 for satellites in near-polar orbits with 400 km altitude.

    - 20 -

  • APPENDIX

    From classical mechanics (Goldstein, 1950), Lagrange’s equation for the motion of a particle is

    given by

    ddt

    ∂T∂qi

    – ∂T∂qi= Qi (A.1)

    where {qi,qi} are generalized coordinates, T is the kinetic energy of the particle, and Qi is a

    component of the generalized force:

    Qi = F j ⋅

    ∂x j∂qi

    Σj

    (A.2)

    F j being the jth force acting on the particle and expressed in inertial Cartesian coordinates:

    {xk} = x = x(qi) ; k = 1,2,3 (A.3)

    The application at hand is the motion of a satellite in orbit around Earth (or any other planet).

    As such the motion is unconstrained in terms of the coordinates and the system is trivially

    holonomic. It is simplest in this case to specialize the generalized coordinates to Cartesian

    coordinates:

    xk = qk (A.4)

    The coordinate frame is assumed to be inertial in the sense of being fixed to Earth’s center of mass

    (it is in free fall in the gravitational fields of the sun, moon, and other planets) and not rotating with

    - 21 -

  • respect to space. Under these premises, the kinetic energy is given by

    T = 12 x2

    (A.5)

    with the further assumption that the satellite has unit mass. The forces acting on the satellite are

    divided into kinematic forces (Martin, 1988) due to the gravitational fields, V, and action forces,

    F , caused variously by atmospheric drag, solar radiation pressure, albedo (Earth-reflected solar

    radiation), occasional thrusting of the satellite as part of orbital maintenance, and a host of other

    minor effects, such as electrostatic and electromagnetic interactions and thermal radiation (Seeber,

    1993). We write for the total force

    F = ∇∇V + F (A.6)

    The total gravitational potential, V, comprises the potentials of all masses in the universe and it is a

    function of position in the inertial frame and of time, but not of velocity:

    V = V(x,t) (A.7)

    We use the sign convention for the potential that is common in geodesy and geophysics. The

    temporal dependence arises from Earth’s rotation (also, not constant); the moon’s, sun’s, and

    planets’ motion relative to the Earth; and the change in potential due to solid Earth tides,

    atmospheric and ocean tides, their loading effects, and other terrestrial mass redistributions of

    secular (e.g., post-glacial rebound) and periodic type.

    Lagrange’s equation derives from the principle of virtual work and ultimately is based on

    Newton’s Second Law of Motion to which one returns upon substituting (A.5) and (A.4) into

    (A.1) and A.2):

    - 22 -

  • ddt

    x = F (A.8)

    where x is also linear momentum (for unit mass). However, equations like (A.1) expressing

    energy relationships are more suited to our purpose since they treat position and momentum as

    distinct coordinates (states) of the system. Along this line, define

    H = T – V (A.9)

    We note that H = H(x,x,t) , and H is the Hamiltonian of the motion only if F = 0 .

    We have

    dHdt

    = ∂H∂xkdxkdtΣk +

    ∂H∂xk

    dxkdtΣk +

    ∂H∂t (A.10)

    Noting the dependencies of T and V on xk , xk , and t, this simplifies to

    dHdt

    = – ∂V∂xkxkΣk +

    dTdxk

    dxkdtΣk –

    ∂V∂t (A.11)

    From (A.6) and (A.8),

    ∂V∂xk

    =dxkdt

    – Fk (A.12)

    and from (A.5), dT dxkdT dxk = xk . Substituting these into (A.11) yields

    dHdt

    = Fk xkΣk –∂V∂t (A.13)

    - 23 -

  • Integrating both sides and using (A.9), we obtain:

    T – V = Fk xk dt

    t0

    t

    Σk

    – ∂V∂t dtt0

    t

    + E0 (A.14)

    where E0 is the constant of integration. If the gravitational potential is static in inertial space

    (principally, no Earth rotation) and if the non-gravitational forces are absent ( F = 0 ), then (A.14)

    expresses the energy conservation law.

    Acknowledgments: The author is grateful to the reviewers for their valuable comments. Thiswork was supported by a grant from the University of Texas, Austin, Contract No. UTA98-0223,under a primary contract with NASA.

    REFERENCES

    Bernard, A. and P. Touboul (1989): A Spaceborne Gravity Gradiometer for the Nineties. Paperpresented at the General Meeting of the International Association of Geodesy, 3-12 August1989, Edinburgh, Scotland.

    Colombo, O.L. (1984): The global mapping of gravity with two satellites. Report of theNetherlands Geodetic Commission, 7(3), Delft.

    Danby, J.M.A. (1988): Fundamentals of Celestial Mechanics. Willman-Bell, Inc., Richmond,Virginia.

    Dickey, J.O. (ed.) (1997): Satellite gravity and the geosphere. Report from the Committee on EarthGravity from Space, National Research Council, National Academy Press.

    Fischell, R.E. and V.L. Pisacane (1978): A drag-free lo-lo satellite system for improved gravityfield measurements. Proceedings of the Ninth GEOP Conference, Report no.280, Departmentof Geodetic Science, Ohio State University, Columbus.

    Goldstein, H. (1950): Classical Mechanics, Addison-Wesley Publ. Co., Reading. Massachusetts.Ilk, K.H. (1986): On the regional mapping of gravitation with two satellites. Proceedings of the

    First Hotine-Marussi Symposium on Mathematical Geodesy, 3-6 June 1985, Rome.Jekeli, C. and R.H. Rapp (1980): Accuracy of the determination of mean anomalies and mean

    - 24 -

  • geoid undulations from a satellite gravity mapping mission. Report no.307, Department ofGeodetic Science, The Ohio State University.

    Jekeli, C. and T.N. Upadhyay (1990): Gravity estimation from STAGE, a satellite-to-satellitetracking mission. Journal of Geophysical Research, 95(B7), 10973-10985.

    Keating, T., P. Taylor, W. Kahn, F. Lerch (1986): Geopotential Research Mission, Science,Engineering, and Program Summary. NASA Tech. Memo. 86240.

    Lambeck, K. (1988): Geophysical Geodesy. Clarendon Press, Oxford.Lemoine, F.G. et al. (1998): The development of the joint NASA GSFC and the National Imagery

    Mapping Agency (NIMA) geopotential model EGM96. NASA Technical Report NASA/TP-1998-206861, Goddard Space Flight Center, Greenbelt, Maryland.

    Martin, J.L. (1988): General Relativity, A Guide to its Consequences for Gravity andCosmology. Ellis Horwood Ltd, Chichester.

    McCarthy, D.D. (1996): IERS Conventions (1996). IERS Technical Note 21, Observatoire deParis, Paris.

    Morgan, S.H. and H.J. Paik (Eds.) (1988): Superconducting Gravity Gradiometer Mission,vol.II, Study Team Technical Report, NASA Tech. Memo. 4091.

    Mueller, I.I. (1969): Spherical and Practical Astronomy. Frederick Ungar Publ. Co., New York.Rummel, R. (1980): Geoid heights, geoid height differences,and mean gravity anomalies from

    “low-low” satellite-to-satellite tracking - an error analysis. Report no.306, Department ofGeodetic Science, Ohio State University, Columbus.

    Rummel, R. and N. Sneeuw (1997): Toward dedicated satellite gravity field missions. Paperpresented at the Scientific Assembly of the IAG, Rio de Janeiro, Brazil, 3-9 September, 1997.

    Seeber, G. (1993): Satellite Geodesy. Walter de Gruyter, Berlin.Seidelmann, P.K. (1992): Explanatory Supplement to the Astronomical Almanac. Prepared by

    U.S. Naval Observatory. University Science Books, Mill Valley, California.Tapley, B.D. and C. Reigber (1998): GRACE: A satellite-to-satellite tracking geopotential mapping

    mission. Proceedings of Second Joint Meeting of the Int. Gravity Commission and the Int.Geoid Commission, 7-12 September 1998, Trieste.

    Torge, W. (1991): Geodesy. Walter de Gruyter, Berlin.Wagner, C.A. (1983): Direct determination of gravitational harmonics from low-low GRAVSAT

    data. Journal of Geophysical Research, 88(B12), 10309-10321.Wolff, M. (1969): Direct measurement of the Earth’s gravitational potential using a satellite pair.

    Journal of Geophysical Research, 74(22), 5295-5300.

    - 25 -