31.07.2013 Views

Accepted Manuscript - Cnr

Accepted Manuscript - Cnr

Accepted Manuscript - Cnr

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Accepted</strong> <strong>Manuscript</strong><br />

Drag and Energy Accommodation Coefficients During Sunspot Maximum<br />

C. Pardini, L. Anselmo, K. Moe, M.M. Moe<br />

PII: S0273-1177(09)00618-8<br />

DOI: 10.1016/j.asr.2009.08.034<br />

Reference: JASR 9965<br />

To appear in: Advances in Space Research<br />

Received Date: 30 January 2009<br />

Revised Date: 22 May 2009<br />

<strong>Accepted</strong> Date: 27 August 2009<br />

Please cite this article as: Pardini, C., Anselmo, L., Moe, K., Moe, M.M., Drag and Energy Accommodation<br />

Coefficients During Sunspot Maximum, Advances in Space Research (2009), doi: 10.1016/j.asr.2009.08.034<br />

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers<br />

we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and<br />

review of the resulting proof before it is published in its final form. Please note that during the production process<br />

errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

Drag and Energy Accommodation Coefficients During Sunspot Maximum<br />

C. Pardini a , L. Anselmo a , K. Moe b and M. M. Moe b<br />

a Space Flight Dynamics Laboratory, ISTI/CNR, Via G. Moruzzi 1, 56124 Pisa, Italy<br />

b Space Environment Technologies, 23 Purple Sage, Irvine, CA 92603, United States<br />

a E-mail: Carmen.Pardini@isti.cnr.it; Luciano.Anselmo@isti.cnr.it<br />

b E-mail : kmmoe2@att.net<br />

Abstract<br />

Conditions appropriate to gas-surface interactions on satellite surfaces in orbit have not been<br />

successfully duplicated in the laboratory. However, measurements by pressure gauges and mass<br />

spectrometers in orbit have revealed enough of the basic physical chemistry that realistic theoretical<br />

models of the gas-surface interaction can now be used to calculate physical drag coefficients. The<br />

dependence of these drag coefficients on conditions in space can be inferred by comparing the<br />

physical drag coefficient of a satellite with a drag coefficient fitted to its observed orbital decay.<br />

This study takes advantage of recent data on spheres and attitude-stabilized satellites to compare<br />

physical drag coefficients with the histories of the orbital decay of several satellites during the<br />

recent sunspot maximum. The orbital decay was obtained by fitting, in a least squares sense, the<br />

semi-major axis decay inferred from the historical two-line elements acquired by the US Space<br />

Surveillance Network. All the principal orbital perturbations were included, namely geopotential<br />

harmonics up to the 16 th degree and order, third body attraction of the Moon and the Sun, direct<br />

solar radiation pressure (with eclipses), and aerodynamic drag, using the Jacchia-Bowman 2006<br />

(JB2006) model to describe the atmospheric density. After adjusting for density model bias, a<br />

comparison of the fitted drag coefficient with the physical drag coefficient has yielded values for<br />

the energy accommodation coefficient as well as for the physical drag coefficient as a function of<br />

altitude during solar maximum conditions. The results are consistent with the altitude and solar<br />

cycle variation of atomic oxygen, which is known to be adsorbed on satellite surfaces, affecting<br />

both the energy accommodation and angular distribution of the reemitted molecules.<br />

1. Introduction<br />

Both accelerometer measurements and orbital decay studies show that satellite drag<br />

coefficient calculations need improving, especially between 300 and 800 km altitude, where many<br />

important satellites fly. Better drag coefficients are needed for improving lifetime predictions of<br />

satellites and for refining thermospheric density models. Atmospheric densities derived from<br />

accelerometer measurements and orbital decay depend on the drag coefficient, CD, assigned to the<br />

satellite, because satellites rarely incorporate a method of measuring it. Consequently, the accuracy<br />

of the inferred densities can be no better than the accuracy of the assumed value of CD. In the past,


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

the use of inaccurate drag coefficients has introduced biases in thermospheric density models<br />

(Chao, et al., 1997; Moe, et al., 2004).<br />

The relationship between the magnitude of the drag force F and atmospheric density ρ is<br />

F = M a = ½ ρ CD A Vi 2 (1)<br />

in which Vi is the speed of the satellite relative to the atmosphere, M is the mass of the satellite, and<br />

A is a reference area, usually taken to be the cross sectional area of the satellite normal to the<br />

incident airstream. The product A Vi 2 is usually well known (except during large geomagnetic<br />

storms). The drag force F is measured either by an accelerometer or by the observed orbital decay.<br />

Therefore, it is the product ρ CD which is determined by the observations.<br />

In practice, three types of satellite drag coefficients have been used: fixed drag coefficients,<br />

fitted drag coefficients, and physical drag coefficients. Fixed drag coefficients simplify the<br />

processes of constructing and using atmospheric density models. Fitted drag coefficients are of<br />

great value in constructing orbits with the aid of an atmospheric model. Physical drag coefficients<br />

are related directly to the drag force, and require a realistic model of the molecular collisions and<br />

adsorptive processes at the satellite surface. They are required for determining absolute atmospheric<br />

densities.<br />

Early in the space age, scientists used a variety of fixed drag coefficients, mostly in the range<br />

of 2.0 to 2.3 (Wolverton, 1963). Outstanding theoretical work was done by Schaaf and Chambré<br />

(1958), Schamberg (1959a and b), and Sentman (1961a and b), but it received little attention from<br />

thermospheric workers at that time. Studies by Izakov (1965) and Cook (1965 and 1966), based on<br />

laboratory measurements, led to the widespread use of a fixed drag coefficient of 2.2 by the<br />

international space community. This had the advantage that there was one less variable when<br />

reducing data or comparing data sets. (The equations for drag and accommodation coefficients<br />

given by Cook are still sometimes used. They agree with idealized laboratory measurements on<br />

clean surfaces, but do not represent the conditions encountered by satellite surfaces in orbit.)<br />

Classified flights of long cylindrical reconnaissance satellites, beginning in 1960, showed that long<br />

cylinders had drag coefficients between 3 and 4, which confirmed the calculations of Sentman<br />

(1961a), but this information was not declassified until 1971 (DeVries, 1971).<br />

Satellite and rocket measurements by pressure (density) gauges (Carter, et al, 1969) and mass<br />

spectrometers (Hedin, et al, 1973; Offermann and Grossmann, 1973) revealed that satellite surfaces<br />

are contaminated by adsorbed molecules. Moe et al. (1972) showed that adsorption on the walls of<br />

Carter’s pressure gauge could be modeled by assuming heterogeneous adsorption, using the<br />

Langmuir isotherm. Hedin, et al., (1973) modeled the adsorption of atomic oxygen in the mass<br />

spectrometer on OGO 6 using the Langmuir isotherm. These modeling studies confirmed that<br />

satellite surfaces are contaminated with adsorbed gases, and emphasized the importance of correctly<br />

representing the energy accommodation and angular distributions of atmospheric molecules<br />

reemitted from satellite surfaces (Moe, et al., 1995, 1998).<br />

In the 1970s, energy accommodation coefficients were measured in low-earth orbit at times of<br />

low and moderate solar activity by paddlewheel satellites (Beletsky, 1970; Imbro, et al. 1975), and<br />

by the ratio of lift to drag on the S3-1 satellite (Ching, et al., 1977). These accommodation<br />

coefficients were 0.99 to 1.00 near 200 km, and about 0.88 near 320 km. Angular distributions of<br />

reemitted molecules have been measured near 180 km by Beletsky (1970), near 225 km by Gregory<br />

and Peters (1987), and near 700 km, by Moe and Bowman (2005). Near 200 km the angular<br />

distribution was about 98% diffuse, while near 700 km it appeared to be nearly diffuse during two<br />

sunspot cycles. These orbital measurements have made it possible to calculate the drag coefficients<br />

of satellites in low Earth orbit up to 300 km with considerable confidence at times of low to<br />

moderate solar activity (Moe, et al., 1995; Moe and Moe, 2005). Above 300 km, and during high


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

solar activity, the uncertainty in these calculations increases rapidly because physical drag<br />

coefficient measurements are so rare. The precise fitted drag coefficients (Willis, et al, 2005) at 800<br />

to 1300 km could help fill this void, and make possible the improvement of thermospheric models<br />

at these high altitudes and during geomagnetic storms. Near sunspot maximum, when the<br />

concentration of atomic oxygen is high, the accommodation coefficients are expected to be higher<br />

and less variable, causing drag coefficients to be lower than at solar minimum (Moe and Moe,<br />

1995). In relation to this, see Figure 7 in Bowman and Moe (2005). The present paper improves the<br />

accommodation coefficients at sunspot maximum given in Moe et al. (1995). A review of sunspot<br />

cycles has recently been published by Vacquero (2007).<br />

In addition to the solar cycle variation of accommodation coefficients, there is the possibility<br />

of a longer-term secular variation: Several recent studies (Marcos et al. 2005; Qian, et al., 2008;<br />

Emmert et al., 2008) have reported a long-term secular decrease in thermospheric temperature and<br />

density caused by the secular increase in atmospheric carbon dioxide. In all of these studies it was<br />

assumed that the drag coefficients of satellites are independent of altitude, solar activity, atomic<br />

oxygen concentration, and other parameters. Since drag coefficients vary with all of these<br />

parameters, it would be of value to investigate the effect of varying drag coefficients on the reported<br />

results. In particular, a secular reduction in atomic oxygen will reduce the accommodation<br />

coefficients and increase the drag coefficients of satellites. The trend in accommodation and drag<br />

coefficients could be measured by flying some paddlewheel satellites, as was done early in the<br />

space age (Reiter and Moe, 1969). Such satellites, measuring both spin and orbital decay, are<br />

capable of determining absolute thermospheric densities as well as drag and accommodation<br />

coefficients. Lacking the advantages of a paddlewheel design, we can try to use satellites of<br />

different shapes to infer the behavior of drag and energy accommodation coefficients. In the present<br />

study, we have taken advantage of the increasing data available to compare theoretical<br />

determinations of satellite drag coefficients with a history of satellite orbital decay during sunspot<br />

maximum. In particular, we have extended earlier research (Bowman and Moe, 2005; Moe and<br />

Bowman, 2005) to altitudes above 500 km, to the maximum of solar cycle 23, and to non-spherical<br />

satellites.<br />

2. Accommodation coefficient<br />

The parameters used in drag coefficient models are the angular distribution of the molecules<br />

reemitted from the satellite surface and the energy accommodation coefficient , defined as<br />

= (Ei – Er)/(Ei – Ew). (2)<br />

Here Ei is the kinetic energy of the incident molecules, Er is the kinetic energy of the reemitted<br />

molecules, and Ew is the energy the reemitted molecules would have if they had adjusted completely<br />

to the surface (or wall) temperature before reemission. In other words, the accommodation<br />

coefficient indicates how closely the kinetic energy of the incoming molecule has adjusted to the<br />

thermal energy of the surface before it is reemitted. If the adjustment is complete, then = 1.00.<br />

This is called complete accommodation. It is important to know the accommodation coefficient,<br />

because CD could be 1.8, or 2.4, or some other value, depending on the accommodation coefficient<br />

and angular distribution. The theoretical determination of drag coefficients requires that we<br />

calculate the momentum transferred to the satellite by the impact of the molecules striking the<br />

surface. That momentum transfer cannot be calculated unless we know the angular distribution and


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

energy loss of the molecules. The energy accommodation coefficient gives us the vital information<br />

about energy loss.<br />

The value of the accommodation coefficient for a particular satellite is dependent on the<br />

stream velocity, Vi , the masses of the incident molecules, and the condition of the surface. Surface<br />

conditions are influenced by a number of parameters. The most important is satellite altitude, which<br />

largely determines the number of oxygen atoms striking the surface and then being adsorbed.<br />

Atomic oxygen binds strongly to many surfaces and changes surface properties (Riley and Giese,<br />

1970; Wood, 1971). It is now well known that satellite surfaces at altitudes of 150-300 km are<br />

contaminated with adsorbed atomic oxygen and its reaction products. The higher the altitude, the<br />

lower the surface fraction covered with adsorbed atoms and molecules. If the incoming molecules<br />

struck a clean surface, they would be reemitted near the specular angle with a partial loss of incident<br />

kinetic energy. But when the surface is contaminated by adsorbed molecules, the incident molecules<br />

lose a large amount of kinetic energy and are reemitted in a diffuse distribution. Adsorbed<br />

molecules then increase energy accommodation and broaden the angular distribution of reflected<br />

molecules. The equation for the drag coefficient of a flat plate, based ultimately on Sentman’s<br />

(1961b) analysis, is given in terms of accommodation coefficients by Moe, et al., (2004). It can be<br />

calculated for other convex shapes by integration.<br />

The brief review of satellite measurements in the introduction has revealed that, in low Earth<br />

orbit near 200 km, is between 1.00 and 0.99 (Moe and Bowman, 2005), and that the angular<br />

distribution of reemitted molecules is diffuse. Sentman’s analysis of gas-surface interactions<br />

(Sentman, 1961b) is then appropriate for calculating the physical drag coefficients of satellites in<br />

the neighborhood of 200 km, because it assumes diffuse reemission. At higher altitudes the<br />

accommodation coefficient decreases, suggesting a lower surface contamination and a quasispecular<br />

reemission of some of the incident molecules. At altitudes near 300 km at solar minimum,<br />

has fallen by about 10%, to the vicinity of 0.9. In this case, Schamberg’s model of drag<br />

coefficients (Schamberg, 1959a and b) could be appropriate for calculating the contribution of the<br />

quasi-specular component, particularly at times of low solar activity. However, an earlier study<br />

(Moe and Bowman, 2005) demonstrated that even at 700 km altitude, the orbital decay of three<br />

Calspheres with different surface materials over a period of 19 years was independent of the surface<br />

material. This suggests that even at 700 km the surfaces are contaminated and the reemission is<br />

mostly diffuse. The present study is made for solar maximum when the most atomic oxygen is<br />

adsorbed on satellite surfaces, so more diffuse reemission is to be expected. Therefore the use of<br />

Sentman’s analysis is justified in the present investigation.<br />

3. Drag and energy accommodation coefficients at sunspot maximum<br />

Ten satellites were involved in this study, including several processed by Bowman and Moe<br />

(2005). Their perigee altitudes were between 200 and 630 km, and their orbital decay was analyzed<br />

during the sunspot maxima of solar cycles 22 and 23. Spherical smooth objects (5 Taifun Yug<br />

satellites and 3 Calspheres) were considered, together with one Surrey satellite (Clementine) and the<br />

Student Nitric Oxide Explorer (SNOE), a small student satellite project of the University of<br />

Colorado. Both Clementine and SNOE had simple shapes and were attitude stabilized in order to<br />

have constant orientation with respect to the airstream. Clementine also slowly rotated about the<br />

vertical axis to prevent overheating. SNOE is pictured in Figure 1, and Clementine in Figure 2.<br />

Table 1 lists the satellites analyzed by Bowman and Moe that decayed during the maximum of<br />

solar cycle 22. Table 2 includes the additional satellites considered in this study during the sunspot


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

maximum of solar cycle 23. The time span covered by our analysis is specified, together with the<br />

perigee/apogee altitude and the orbital inclination of each satellite at the beginning of the time span.<br />

Drag coefficients for the various shapes of the bodies in the study (see Tables 1 and 2) were<br />

calculated using a range of energy accommodation coefficients. This was carried out by applying<br />

Sentman’s analysis of the interaction of incoming particles with a satellite surface to determine the<br />

component of the force along the satellite's direction of motion. Sentman’s assumption of diffuse<br />

reemission of the particles is appropriate as discussed in Section 2. In addition to the<br />

accommodation coefficient, , the input quantities used for calculating the drag coefficients<br />

included satellite speed and the temperature and mean molecular mass of the ambient atmosphere,<br />

together with the parameters related to the shape of the satellite. For each satellite, a series of drag<br />

coefficient values was obtained as a function of . The drag coefficients so determined are named<br />

“physical drag coefficients”, to distinguish them from drag coefficients determined by fitting the<br />

orbital decay to a particular density model.<br />

The “fitted drag coefficients” were computed by an analysis of the orbital decay of satellites.<br />

In some cases the drag coefficients so obtained were updated, when more recent information<br />

regarding the dimensions of the satellite became available. The fitted drag coefficients were then<br />

adjusted upward by a certain amount (dependent on altitude) to account for the known biases in the<br />

atmospheric density model used at sunspot maximum. The modified fitted drag coefficients are<br />

called “observed drag coefficients”.<br />

After having found the physical drag coefficient that matched the observed drag coefficient,<br />

the corresponding energy accommodation coefficient was identified.<br />

3.1 Fitted drag coefficients<br />

For each satellite, the fitted drag coefficient was obtained by fitting, in a least squares sense,<br />

the semi-major axis decay inferred from the historical two-line elements (TLEs) provided by the US<br />

Space Surveillance Network (Hoots and Roehrich, 1980) . The inverse ballistic coefficient of the<br />

satellite is defined as<br />

B = CD A/M (3)<br />

where CD, A and M are, respectively, the satellite drag coefficient, cross-sectional area and mass.<br />

In orbit fitting, B was adjusted to force the atmospheric density model to agree with the air drag<br />

revealed by the tracking data, i.e. the historical TLE record. Having fixed the area-to-mass ratio<br />

A/M, the only solve-for parameter of the fit, in the least squares sense, was the satellite drag<br />

coefficient CD. The semi-major axis root mean square (rms) residuals (R) were computed according<br />

to the relationship (Pardini and Anselmo, 2008)<br />

N<br />

<br />

<br />

a a <br />

i _ obs<br />

i _ com<br />

2<br />

i 1<br />

R (4)<br />

N<br />

where ai_obs and ai_com are, respectively, the observed and the computed semi-major axis at the<br />

same epoch and N is the number of observations available, i.e. the number of TLEs used in the


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

fitting.<br />

The software code CDFIT, specifically developed at ISTI/CNR and adopting the same force<br />

model as the orbit propagator SATRAP (Pardini and Anselmo, 1994), was used to fit the semimajor<br />

axis decay. All the main orbital perturbations were considered, namely geopotential<br />

harmonics up to the 16 th degree and order (EGM96 model, Lemoine et al., 1998), third body<br />

attraction of the Moon and the Sun, direct solar radiation pressure with eclipses, and aerodynamic<br />

drag.<br />

The recent Jacchia-Bowman 2006 (JB2006) model (Bowman et al., 2008) was used to<br />

describe the atmospheric density. (Several other empirical density models, such as NRLMSISE-00<br />

(Picone et al, 2002), DTM-2000 (Bruinsma et al., 2003), GOST-2004 (Volkov, 2004) are widely<br />

used, but we chose JB2006 because its biases had been measured.) The JB2006 indices (Tobiska et<br />

al., 2008) were those released by Space Environment Technologies (SET, 2008) on 13 June 2007<br />

(release v3.8). Figure 3 shows the 81-day centered smoothed values of the v3.8 JB2006 solar<br />

indices F10.7 (F81c), S10.7 (S81c) and M10.7 (M81c) during the sunspot maximum of solar cycle 23 (1<br />

October 1999 – 31 December 2002).<br />

3.1.1 Impact of mismodeled radiation pressure and other minor perturbations on fitted drag<br />

coefficients<br />

For satellites above 500 km, a mismodeling of radiation pressure, namely direct solar<br />

radiation pressure, Earth-reflected radiation and Earth-emitted infrared radiation, might critically<br />

affect the accuracy of the fitted drag coefficients. To assess the impact of such possible<br />

miscalculations on our fitted drag coefficients and considering that they were determined from<br />

semi-major axis decay, the orbit evolution of two satellites (one in the neighborhood of 500 km,<br />

SNOE, and the other near 650 km, Clementine), was analyzed by disregarding or including the<br />

various perturbations caused by radiation pressure. The special perturbations trajectory propagator<br />

SATORB (Beutler, 2005) was used to carry out this analysis.<br />

The orbit of SNOE (initial perigee/apogee altitude = 515/570 km, inclination = 97.7, A/M =<br />

0.007 m 2 /kg) was propagated for 60 days (i.e. two times the typical length of the arcs used to obtain<br />

fitted drag coefficients), including 16 16 geopotential, luni-solar attraction and air drag, plus<br />

various combinations of smaller perturbations. The overall mean semi-major axis decay, due<br />

basically to air drag, was about 6 km. The total exclusion of direct solar radiation pressure resulted<br />

in an average mean semi-major axis error of 0.729 m. The average mean semi-major axis error<br />

introduced by disregarding the Earth’s albedo was 0.066 m. It was also found that the<br />

corresponding mean errors were 0.019 m for neglecting the Earth’s tides and 0.001 m for neglecting<br />

the relativistic effects.<br />

The orbit evolution of Clementine (initial perigee/apogee altitude = 643/663 km, inclination =<br />

98.1, A/M = 0.011 m 2 /kg) was simulated for 3 years (i.e. the length of the greatest arc used in this<br />

case to obtain the fitted drag coefficient), again with 16 16 geopotential, luni-solar attraction, air<br />

drag in conditions of high solar activity, and various combinations of radiation pressure and smaller<br />

perturbations. The overall mean semi-major axis decay, due basically to air drag, was about 47 km.<br />

The total exclusion of direct solar radiation pressure resulted in an average mean semi-major axis<br />

error of 16.440 m. A mismodeling of direct solar radiation pressure by 20% resulted in an average<br />

mean semi-major axis error of 3.037 m. The average mean semi-major axis error introduced by<br />

disregarding the Earth’s albedo was 1.406 m. It was also found that the corresponding mean errors<br />

were 0.054 m for neglecting the Earth’s tides and 0.010 m for neglecting the relativistic effects.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

In conclusion, the effects on the mean semi-major axis of Earth’s albedo (reflected and<br />

thermal emitted radiation) and other smaller perturbations were less than those due to direct solar<br />

radiation pressure mismodeling. In addition, even direct solar radiation pressure with eclipses<br />

(included in the force model used for drag coefficient fits) had a very small influence on mean<br />

semi-major axis evolution and dispersion in the altitude regimes considered.<br />

As a further test, fitted drag coefficients were computed with CDFIT for the Optical<br />

Calibration Sphere (OCS) satellite (initial perigee/apogee altitude = 748/803 km, inclination =<br />

100.2), varying solar radiation pressure by as much as 100%. However, even for this large area-tomass<br />

ratio satellite (A/M = 0.543 m 2 /kg), the maximum difference of the drag coefficient estimate<br />

was 0.5% or less. Finally, a mismodeling of direct solar radiation pressure by 20% had no<br />

detectable effect (i.e. 0.01) on the CD estimation for the satellites Clementine and UOSAT-2<br />

(initial perigee/apogee altitude = 648/663 km, inclination = 97.9, A/M = 0.007 m 2 /kg).<br />

3.2 SNOE drag and energy accommodation coefficients<br />

SNOE had a cylinder-like shape with a hexagonal cross section. It was attitude stabilized, so<br />

that it maintained a constant aspect relative to the incident velocity vector, a feature that facilitated<br />

the computation of its drag coefficient as a function of . After its first year, the attitude sometimes<br />

deviated from the nominal by ten degrees. SNOE was in a nearly circular orbit with a perigee<br />

altitude of 515 km at the beginning of our analysis (Figure 4). Its semi-major axis was observed to<br />

decay by about 38 km per year from 1 October 1999 to 31 December 2002 (Figure 4). In the same<br />

time interval, 40 values of the SNOE fitted drag coefficients were obtained: the first 39 over 30-day<br />

intervals, the last over a 15-day interval. Figure 5 shows the fitted CD values, together with the<br />

corresponding inverse ballistic coefficient (B) and semi-major axis root mean square (rms)<br />

residuals, defined according to Eqs. 3 and 4, respectively. A cross-sectional area of 1 m 2 and a mass<br />

of 115.5 kg were assumed in order to estimate the fitted drag coefficients. The average fitted CD<br />

was found to be 1.810. The standard deviation of an individual measurement (), defined according<br />

to<br />

<br />

N<br />

<br />

n1<br />

2<br />

xn x<br />

N 1<br />

where xn represents the n th measurement and x is the arithmetic mean of all N measurements, was<br />

0.184. If each of the 40 monthly measurements were statistically independent, the standard<br />

deviation of the average, i.e. / N , with N = 40, would be 0.029. However, there are about four<br />

positive peaks in the 40 month record. This implies that the 40 monthly samples are not<br />

independent. We infer that the autocorrelation function has eight zeros (Rice, 1954); in other<br />

words, there are only eight independent samples. Therefore, / N<br />

is 0.065.<br />

Two adjustments were then made to calculate the observed drag coefficient of SNOE. First,<br />

the average area facing the air stream, according to the exact dimensions of SNOE, was 0.8377 m 2 ,<br />

compared to the approximate value of 1 m 2 . The average drag coefficient (fitted to JB2006) then<br />

became 2.16. This value was later increased by 15% to correct for the bias in the JB2006 density<br />

model at an altitude of around 500 km (see Table 3 for the biases of the modified Jacchia 1970<br />

(5)


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

model (J70 MOD), obtained by Bowman and Moe (2005), and the biases of JB2006 obtained by us<br />

with Bowman’s help). The resulting observed drag coefficient of SNOE was therefore 2.485.<br />

A comparison of this result with the calculated dependence of the SNOE physical drag<br />

coefficient on the energy accommodation coefficient revealed an agreement for = 0.95. This high<br />

accommodation coefficient does not fit smoothly in the sequence, perhaps because the satellite<br />

orientation sometimes deviated by ten degrees from the nominal. at sunspot minimum from Moe<br />

et al. (1995) is given in Table 6 for comparison. Near 200 km, is 1.00 at sunspot maximum and<br />

0.99 at sunspot minimum; but at sunspot maximum, decreases more slowly as the altitude<br />

increases above 200 km. This is physically reasonable, because during sunspot maximum the<br />

increased solar UV radiation dissociates more molecular oxygen, and increased UV heating causes<br />

the resulting atomic oxygen to rise to higher altitudes, increasing the energy accommodation.<br />

3.3 Clementine drag and energy accommodation coefficients<br />

Clementine was a gravity-gradient stabilized rectangular box (0.60 x 0.35 x 0.35 m 3 ), with<br />

four solar panels (0.53 x 0.35 m 2 ) tilted backward at 45° (relative to the normal to the box surface),<br />

plus a 6 meter long rod for gravity-gradient stabilization. The long axis of the spacecraft pointed<br />

within 1° of the nadir direction. The vehicle slowly rotated about its long axis to prevent any<br />

overheating of the instruments within the satellite. With an initial perigee altitude in the vicinity of<br />

650 km (Figure 6), the satellite did not experience a strong drag perturbation. The observed mean<br />

semi-major axis decay was about 11 km per year over the time span covered by our analysis, i.e.<br />

from 3 December 1999 to 31 December 2002 (Figure 6). The effects of air drag on semi-major axis<br />

then became significant only over a relatively long period, due to the secular nature of the induced<br />

decay. In this case one-month arcs were too short to obtain sufficiently good fits of the drag<br />

coefficient. Therefore, CD was estimated over orbital arcs of 180, 360 and 1080 days, over the time<br />

interval of 8 December 1999 to 21 November 2002. The average fitted drag coefficients were 2.63<br />

(over 6 180-day arcs), 2.65 (over 3 360-day arcs) and 2.64 (over a 1080-day arc) for an assumed<br />

reference area of 0.48 m 2 and a mass of 50 kg. However, the actual cross-sectional area of the<br />

satellite encountered by the airstream was computed to be 0.568 m 2 . This value included the<br />

projected area of the central box, the solar panels and the long rod. By then applying the ratio<br />

0.48/0.568 to adjust the reference area and the correction for the bias in the JB2006 model (15% at<br />

630 km; see Table 3), the resulting observed drag coefficients were 2.56, 2.58 and 2.57,<br />

respectively.<br />

The physical drag coefficient was calculated by determining the momentum transfer to the<br />

satellite caused by the incident air molecules and the molecules reemitted from the surface in a<br />

diffuse angular distribution, with energy loss given by the energy accommodation coefficient. Since<br />

the surfaces were rotating flat plates, the resultant drag force was determined by averaging over the<br />

angles of rotation. The other input parameters used included a satellite velocity of 7,540 m/s, an<br />

ambient atmospheric temperature of 1,260 K, and a mean molecular mass of 12 amu. The results of<br />

the calculations are given in Table 4 as a function of a range of accommodation coefficients.<br />

The next step was to compare the physical drag coefficients in Table 4 with the observed drag<br />

coefficients derived from fitting the orbit, i.e. with CD = 2.56, 2.58 and 2.57. By means of the linear<br />

interpolation in Table 4, Clementine’s average observed CD of 2.57 corresponds with an<br />

accommodation coefficient of 0.875.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

3.4 Cosmos 2265 and Cosmos 2332 drag and energy accommodation coefficients<br />

Both Cosmos 2265 and Cosmos 2332 were spheres with a smooth surface. Their crosssectional<br />

area and mass were assumed to be 3.142 m 2 and 750 kg, respectively. They had an<br />

elliptical orbit with initial perigee altitude in the neighborhood of 279 km for Cosmos 2265 (Figure<br />

7) and 286 km for Cosmos 2332 (Figure 8). Their orbit inclination was around 83.<br />

For Cosmos 2265, the observed semi-major axis decay was nearly 109 km per year and fitted<br />

drag coefficients were obtained over 30-day intervals, from 1 October 1999 to 14 January 2003.<br />

Figure 9 shows the fitted CD values of Cosmos 2265, together with the corresponding inverse<br />

ballistic coefficient (B) and semi-major axis root mean square (rms) residuals. The average fitted<br />

drag coefficient of Cosmos 2265 was 1.869. Its standard deviation was 0.152, and the standard<br />

deviation of the mean of the 40 measurements would have been 0.024 if all 40 measurements had<br />

been independent. Because only eight of the measurements were independent, the standard<br />

deviation of the mean was 0.054.<br />

For Cosmos 2332, the observed semi-major axis decay was nearly 89 km per year and fitted<br />

drag coefficients were obtained over 30-day intervals, from 1 October 1999 to 14 January 2003.<br />

Figure 10 shows the fitted CD values of Cosmos 2332, together with the corresponding inverse<br />

ballistic coefficient (B) and semi-major axis root mean square (rms) residuals. The average fitted<br />

drag coefficient of Cosmos 2332 was 1.920. Its standard deviation was 0.247, and the standard<br />

deviation of the mean of the 40 measurements would have been 0.039 if they all were independent;<br />

but it was 0.087 because there were only eight independent measurements.<br />

The fitted drag coefficients of the two smooth Taifun-1 Yug spheres (Table 2) were below<br />

2.0. This confirmed that the JB2006 density model was biased at sunspot maximum. Therefore, by<br />

applying the correction for the bias in the density model (Table 3) at an average altitude of 275 km,<br />

the observed drag coefficients were 2.08 and 2.14 for Cosmos 2265 and Cosmos 2332, respectively.<br />

The corresponding accommodation coefficients were 1.00 and 0.993.<br />

3.5 Overall results<br />

Similar calculations were performed for the other satellites processed by Bowman and Moe,<br />

(2005). Bowman and Moe used the modified Jacchia 1970 thermospheric model, for which the<br />

biases also are given in Table 3. The drag and accommodation coefficients at sunspot maximum<br />

derived in the present study are collected in Table 5. For comparison, the accommodation<br />

coefficients at sunspot minimum from Moe, et al. (1995) are tabulated in Table 6. At both sunspot<br />

maximum and minimum, energy accommodation was complete near 200 km, and decreased as the<br />

altitude increased. But accommodation decreased faster at solar minimum. This was expected,<br />

because atomic oxygen controls accommodation coefficients, and there is less atomic oxygen and a<br />

lower temperature at solar minimum.<br />

The relationships of the various drag coefficients and accommodation coefficients in Table 5<br />

(for sunspot maximum) and Table 6 (for sunspot minimum) give an idea of the accuracy: Perhaps<br />

3% for the drag coefficients of spheres and a higher uncertainty for more complicated shapes. An<br />

independent estimate of the accuracy of drag coefficient calculations can be ascertained by<br />

considering the following: A similar Sentman-based analysis of the completely independent US Air<br />

Force data base of accelerometer measurements from three compact Atmosphere Explorer satellites


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

and three long cylindrical satellites reduced the average discrepancy between measurements and<br />

models at 200 km from 9% to 3% (Moe et al., 1998). We attribute this 3% to errors in calculating<br />

the average drag coefficients of spheres and cylinders at 200 km.<br />

As Table 5 and Table 6 show, the relationship between accommodation coefficients and<br />

drag coefficients is nonlinear. It also is different for different shapes. Relationships for several<br />

simple shapes were given in Geophysical Monograph 87 (Moe et al., 1995). One purpose of the<br />

present paper is to improve the accommodations coefficients given in that earlier work.<br />

4. Lessons learned from the analysis of the orbital decay of other satellites<br />

Many other satellites were processed in this study, in order to improve the reliability of the<br />

results obtained. The information from a number of objects was incomplete or had internal<br />

inconsistencies that prevented us from keeping them in the final collection of satellites.<br />

Nonetheless, meaningful lessons were learned from their orbital decay analyses. Here are two<br />

examples:<br />

The orbital decay of seven Taifun-1 Vektor satellites, with solar cells on their spherical<br />

surfaces, was analyzed with the aim of reducing the uncertainty in the drag and accommodation<br />

coefficients in the vicinity of 400 km (close to the perigee altitude of these satellites). Fitted drag<br />

coefficients were calculated over 150-day intervals, between 1 October 1999 and 14 January 2003<br />

(i.e. at sunspot maximum of solar cycle 23). The mean fitted drag coefficient for the seven Vektors,<br />

at an average altitude near 400 km, was 2.591, with a standard deviation of the mean of the seven<br />

measurements of 0.023. These data provided a quantitative indication of the increased drag<br />

coefficients of “nearly spherical” satellites, which is attributed to various deviations from sphericity.<br />

Previous studies have already shown this effect (Bowman and Moe, 2005; Moe and Bowman,<br />

2005), but not with such precision. A recent study by Pilinski (2009), using precise engineering<br />

drawings of Starshines 1 and 2, demonstrated that the actual shapes and cross sectional areas were<br />

different from those publicized: The satellites had retained their attachment rings in orbit.<br />

Some anomalies observed in the fitted drag coefficients of the Optical Calibration Sphere,<br />

OCS, led to additional investigations. It was learned that the satellite, launched in a small canister<br />

and expanded in orbit into a sphere with a diameter of 3.5 m, remained attached to the canister<br />

(Guidamean, 2009). This would explain the OCS drag coefficient, which appeared to be different<br />

from that of a smooth sphere, and to lie between the drag coefficients of smooth spheres and<br />

modified spheres such as the Starshines and the Taifun-1 Vektors. It now appears that the attached<br />

canister increased the average area and changed the shape, so that the drag coefficient was not<br />

actually measured. These new results confirm that satellite dimensions in orbit often are<br />

misreported, and place in doubt the drag coefficients derived for all of the modified spherical<br />

satellites.<br />

5. Conclusions and future work<br />

A detailed analysis of ten satellites with smooth surfaces in low Earth orbit has revealed that:


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

Energy accommodation coefficients are higher at sunspot maximum than at minimum, and<br />

decrease more slowly with increasing altitude. These results are consistent with the larger<br />

amount of atomic oxygen that is adsorbed on satellite surfaces at solar maximum;<br />

Consequently, incident molecules rebound with less energy, so satellite drag coefficients<br />

are generally lower at solar maximum than at minimum;<br />

When spherical satellites are modified by attaching solar cells and other objects, the area<br />

and shape are changed and often misreported, making it difficult to calculate an accurate<br />

drag coefficient.<br />

Knowledge of drag coefficients can be improved by further measurements of accommodation<br />

coefficients using additional non-spherical, attitude-controlled satellites with perigee altitudes above<br />

300 km. The effects of high orbital eccentricities, e.g. geosynchronous transfer orbits, on<br />

accommodation coefficients should also be studied. Furthermore, the reported thermospheric<br />

cooling by carbon dioxide, with its secular density decrease, implies a secular change in drag and<br />

accommodation coefficients. This effect should be investigated. Paddlewheel satellites, similar to<br />

those by means of which the first orbital measurements of energy accommodation were made, could<br />

also be used for these purposes.<br />

Acknowledgments<br />

The results described in this paper were presented at the 37 th COSPAR Scientific Assembly,<br />

held in Montréal, Canada, in July 2008. Anselmo and Pardini contributed to the study within the<br />

framework of the ASI/CISAS Contract No. I/046/07/0.<br />

We would like to thank Stan Solomon and Scott Bailey for information on the SNOE satellite;<br />

Alex da Silva Curiel for information on the Surrey satellites, Clementine, UoSat 2, and Tzinghua-1;<br />

Martin Sweeting and Audrey Nice for providing us with the drawing of Clementine; Koorosh<br />

Guidamean for information on the Optical Calibration Sphere; the US Space Surveillance Network<br />

for making available the TLEs of the satellites; and Space Environment Technologies, for providing<br />

the historical solar indices used in the JB2006 model.<br />

The editor, Pascal Willis, and the reviewers, including Bruce Bowman, made many valuable<br />

suggestions for improving this paper.<br />

References<br />

Beletsky, V. V. An estimate of the character of the interaction between the airstream and a satellite,<br />

Kosmicheskie Issledovaniya 8 (2), 206-217 (in Russian), 1970.<br />

Beutler, G. Methods of Celestial Mechanics, Part. II: Application to Planetary System,<br />

Geodynamics and Satellite Geodesy, Astronomy and Astrophysics Library, Springer-Verlag,<br />

Berlin and Heidelberg, Germany, 2005.<br />

Bowman, B. R., Moe, K. Drag coefficient variability at 175-500 km from the orbit decay analyses<br />

of spheres, Paper AAS 05-257, Presented at the 2005 AAS/AIAA Astrodynamics Conference,<br />

South Lake Tahoe, California, USA, 2005.<br />

Bowman, B. R., Tobiska, W. K., Marcos, F. A., Vallares, C. The JB2006 empirical thermospheric<br />

density model, Journal of Atmospheric and Solar-Terrestrial Physics 70 (5), 774-793, 2008.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

Bruinsma, S., Thuillier, G., Barlier, F. The DTM-2000 empirical thermospheric model with new<br />

assimilation and constraints at lower boundary, accuracy and properties. J. Atmos Solar<br />

Terrestr. Phys. 65, 1053-1070, 2003.<br />

Carter, V. L., Ching, B. K., Elliott, D. D. Atmospheric density above 158 kilometers, J. Geophys.<br />

Res. 74, 5083-5091, 1969.<br />

Chao, C. C., Gunning, G. R., Moe, K., Chastain, S. H., Settecerri, T. J. An evaluation of Jacchia 71<br />

and MSIS90 atmosphere models with NASA ODERACS decay data, J. Astronautical Sciences<br />

45, 131-141, 1997.<br />

Ching, B. K., Hickman, D. R., Straus, J. M. Effects of atmospheric winds and aerodynamic lift on<br />

the inclination of the orbit of the S3-1 Satellite. J. Geophys. Res. 82, 1474-1480, 1977.<br />

Cook, G. E. Satellite drag coefficients, Planetary and Space Science 13, 929-946, 1965.<br />

Cook, G. E. Drag coefficients of spherical satellites, Annales of Geophysique 22, 53-64, 1966.<br />

DeVries, L. L. Private communication, 1971.<br />

Emmert, J. T., Picone, J. M., Meier, R. R. Thermospheric global average density trends, 1967-2007,<br />

derived from orbits of 5000 near-earth objects, Geophys. Res. Lett., 35 L05101,<br />

doi:10.1029/2007GL032809, 2008.<br />

Gregory, J. C., Peters, P. N. A measurement of the angular distribution of 5 eV atomic oxygen<br />

scattered off a solid surface in Earth orbit. Proceedings of the 15th International Symposium on<br />

Rarefied Gas Dynamics, Vol. 2, B. G. Teubner, Stuttgart, Germany, pp. 644-654, 1987.<br />

Guidamean, Koorosh, Personal communication, 2009.<br />

Hedin, A. E., Hinton, B. B., Schmitt, G. A. Role of gas-surface interactions in the reduction of<br />

OGO 6 neutral particle mass spectrometer data, J. Geophys. Res. 78, 4651-4668. 1973.<br />

Hoots, F.R., Roehrich, R.L. Models for Propagation of NORAD Elements Sets, Spacecraft Report<br />

No. 3, Project Spacetrack, Aerospace Defence Command, United States Air Force, Colorado<br />

Springs, Colorado, USA, December 1980.<br />

Imbro, D.R., Moe, M. M., Moe, K. On fundamental problems in the deduction of atmospheric<br />

densities from satellite drag, J. Geophys. Res. 80, 3077-3086, 1975.<br />

Izakov, M. N. Some problems of investigating the structure of the upper atmosphere and<br />

constructing its models, Space Res. V., 1191, 1965.<br />

Lemoine, F. G., Kenyon, S. C., Factor, J. K., et al. The development of the joint NASA GSFC and<br />

NIMA geopotential model EGM96, NASA/TP-1998-206861, NASA Goddard Space Flight<br />

Center, Greenbelt, MD, USA, July 1998.<br />

Marcos, F. A., Wise, J. O., Kendra, M. J., Grossbard, N. J.,Bowman, B. R. Detection of a long-term<br />

decrease in thermospheric neutral density, Geophys. Res. Let. 32, L04103,<br />

doi:1029/2004GL021269, 2005.<br />

Moe, K., Moe, M. M., Yelaca, N. W. Effect of surface heterogeneity on the adsorptive behavior of<br />

orbiting pressure gages, J. Geophys. Res. 77, 4242-4247, 1972.<br />

Moe, M. M., Wallace, S. D., Moe, K. Recommended drag coefficient for aeronomic satellites, The<br />

Upper Mesosphere and Lower Thermosphere: A Review of Experiment and Theory,<br />

Geophysical Monograph 87, 349-356, 1995.<br />

Moe, K., Moe, M. M., Wallace, S. D. Improved satellite drag coefficient calculations from orbital<br />

measurements of energy accommodation, Journal of Spacecraft and Rockets 35 (3), 266-272,<br />

1998.<br />

Moe, K., Moe, M. M., Rice, C. J. Simultaneous analysis of multi-instrument satellite measurements<br />

of atmospheric density, J. Spacecraft and Rockets 41, 849-853, 2004.<br />

Moe, K., Bowman, B. R. The effects of surface composition and treatment on drag coefficients of<br />

spherical satellites, Paper AAS 05-258, Presented at the 2005 AAS/AIAA Astrodynamics<br />

Conference, South Lake Tahoe, California, USA, 2005.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

ACCEPTED MANUSCRIPT<br />

Moe, K., Moe, M. M. Gas-surface interactions and satellite drag coefficients, Planetary and Space<br />

Sciences 53, 793-801, 2005.<br />

Offermann, D., Grossmann, K. U. Thermospheric density and composition as determined by a mass<br />

spectrometer with cryo ion source, J. Geophys. Res. 78, 8296-8304, 1973.<br />

Pardini, C., Anselmo, L. SATRAP: Satellite Reentry Analysis Program, Internal Report C94-17,<br />

Istituto CNUCE, CNR, Pisa, 30 Agosto 1994.<br />

Pardini, C., Anselmo, L. Impact of the time span selected to calibrate the ballistic parameter on<br />

spacecraft re-entry predictions, Advances in Space Research 41 (7), 1100-1114, 2008.<br />

Picone, J. M., Hedin, A. E., Drob, D. P., Aikan, A.C. NRLMSISE-00 empirical model of the<br />

atmosphere: statistical comparisons and scientific issues, Journal of Geophysical Research 107<br />

(A12), 1468, doi:10.1029/2002JA009430, 2002.<br />

Pilinski, M., personal communication, 2009.<br />

Qian, L., Solomon, S. C., Kane, J. K. Seasonal variation of thermospheric density and composition,<br />

Journal of Geophysical Research, in press, 2008.<br />

Reiter, G. S., Moe, K. Surface-particle interaction measurements using paddlewheel satellites, Sixth<br />

International Symposium on Rarefied Gas Dynamics, Vol. 2, 1543-1555, 1969.<br />

Rice, S. O. Mathematical analysis of random noise, in Wax, Nelson. Selected Papers on Noise and<br />

Stochastic Processes, Dover Publications, NY, 1954.<br />

Riley, J. A., Giese, C. F. Interaction of atomic oxygen with various surfaces, Journal of Chemical<br />

Physics 53, 146-152, 1970.<br />

Schaaf, S. A., Chambré, P. L. The flow of rarefied gases, Section H of Fundamentals of Gas<br />

Dynamics, ed. H. W. Emmons, Princeton, New Jersey, USA, pp. 687-739, 1958.<br />

Schamberg, R. A new analytic representation of surface interaction for hyperthermal free molecule<br />

flow with application to neutral-particle drag estimates of satellites, RM-2313 Rand Corp.,<br />

Santa Monica, CA, USA, 1959a.<br />

Schamberg, R. Analytic representation of surface interaction for free molecule flow with<br />

application to drag of various bodies, in Masson, D. J., R-339, Aerodynamics of the Upper<br />

Atmosphere, pp. 12-1 to 12-41, Rand Corp., Santa Monica, CA, USA, 1959b.<br />

Sentman, L. H. Comparison of the exact and approximate methods for predicting free molecule<br />

aerodynamic coefficients. American Rocket Soc. J. 31, 1576-1579, 1961a.<br />

Sentman, L. H. Free molecule flow theory and its application to the determination of aerodynamic<br />

forces, Lockheed Missile and Space Co., LMSC-448514, AD 265-409, Sunnyvale, CA, USA,<br />

1961b.<br />

SET, 2008. Space Environment Technologies: http://www.spacewx.com/.<br />

Tobiska, W. K., Bouwer, S. D., Bowman, B. R. The development of new solar indices for use in<br />

thermospheric density modeling, Journal of Atmospheric and Solar-Terrestrial Physics 70 (5),<br />

803-819, 2008.<br />

Vaquero, J. M. Historical sunspot observations, a review, Adv. Space Res., 40 (7), 929-941, 2007.<br />

Volkov, I. I. Earth’s upper atmosphere density model for ballistic support of the flight of artificial<br />

Earth satellites GOST R 25645.166-2004. Publishing House of the Standards, Moscow, 2004.<br />

Willis, P., Deleflie, F., Barlier, F., et al. Effects of thermosphere total density perturbation on LEO<br />

orbits during severe geomagnetic conditions (Oct - Nov 2003) using DORIS and SLR data,<br />

Adv. Space Res., 36(3), 522-533, 2005.<br />

Wolverton, R. W. Flight Performance Handbook for Orbital Operations. Wiley, N.Y. and London,<br />

pp. 2-336 – 2-339, 1963.<br />

Wood, B. J. The rate and mechanism of interaction of oxygen atoms and hydrogen atoms with<br />

silver and gold, Journal of Physical Chemistry 75, 2186-2193, 1971.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

Figure captions<br />

ACCEPTED MANUSCRIPT<br />

Figure 1. Graphical representation of the Student Nitric Oxide Explorer (SNOE) satellite. Courtesy<br />

of Stan Solomon of University Corporation for Atmospheric Research (UCAR).<br />

Figure 2. Graphical representation of the Surrey satellite Clementine. Courtesy of Martin Sweeting<br />

and Audrey Nice of Surrey Satellite Technologies LTD.<br />

Figure 3. 81-day centered smoothed values of the v3.8 JB2006 solar indices F10.7 (F81c), S10.7<br />

(S81c) and M10.7 (M81c) during the sunspot maximum of solar cycle 23 (1 October 1999 – 31<br />

December 2002).<br />

Figure 4. SNOE observed perigee/apogee altitude and mean semi-major axis decay from October<br />

1999 to December 2002.<br />

Figure 5. Fitted drag coefficients, inverse ballistic coefficients and rms residuals on semi-major<br />

axis for SNOE.<br />

Figure 6. Clementine observed perigee/apogee altitude and mean semi-major axis decay from<br />

December 1999 to December 2002.<br />

Figure 7. Cosmos 2265 observed perigee/apogee altitude and mean semi-major axis decay from<br />

October 1999 to December 2002.<br />

Figure 8. Cosmos 2332 observed perigee/apogee altitude and mean semi-major axis decay from<br />

October 1999 to December 2002.<br />

Figure 9. Fitted drag coefficients, inverse ballistic coefficients and rms residuals on semi-major<br />

axis for Cosmos 2265.<br />

Figure 10. Fitted drag coefficients, inverse ballistic coefficients and rms residuals on semi-major<br />

axis for Cosmos 2332.


Satellite<br />

Name<br />

ACCEPTED MANUSCRIPT<br />

Table 1<br />

Sample of satellites analyzed by<br />

Bowman and Moe (2005) during the maximum of solar cycle 22 (1989-1990)<br />

Catalog<br />

Number<br />

Diameter<br />

[m]<br />

Mass<br />

[kg]<br />

Sphere<br />

Surface<br />

Inclination<br />

[deg]<br />

Launch<br />

Height<br />

[km]<br />

Launch<br />

Date<br />

Decay<br />

Date<br />

CALSPHERES<br />

Calsphere 3 4957 0.254 0.73 Aluminum 88.3 775 17-Feb-71 17-Oct-89<br />

Calsphere 4 4958 0.254 0.73 Aluminum 88.3 775 17-Feb-71 20-Sep-89<br />

Calsphere 5 4963 0.254 0.73 Gold 88.3 775 17-Feb-71 7-Jan-90<br />

Cosmos 1179 11796 2.000 75010<br />

TAIFUN YUGS<br />

Smooth 82.9 300 *<br />

14-May-80 18-Jul-89<br />

Cosmos 1427 13750 2.000 75010 Smooth 65.8 450 29-Dec-82 5-Oct-89<br />

Cosmos 1615 15446 2.000 75010 Smooth 65.8 450 20-Dec-84 15-Apr-90<br />

* The orbit of Cosmos 1179 was elliptical, the launch height refers to the perigee


ACCEPTED MANUSCRIPT<br />

Table 2<br />

Sample of satellites analyzed<br />

during the sunspot maximum of solar cycle 23 (1999-2002)<br />

Satellite Catalog Area<br />

Name Number [m 2 Mean Orbital Elements<br />

Time span<br />

Mass<br />

at the beginning of the time span over which<br />

] [kg] Inclination Perigee Apogee the orbital<br />

[deg] Altitude Altitude decay was<br />

[km] [km] analyzed<br />

Cylinder-like body with a hexagonal cross section<br />

SNOE 25233 0.8377 115.5 97.7 515 570 1-Oct-99<br />

31-Dec-02<br />

Spinning box with solar panels and a long rod for gravity-gradient stabilization<br />

Clementine 25968 0.568 50 98.1<br />

TAIFUN YUGS<br />

643 663 8-Dec-99<br />

31-Dec-02<br />

Cosmos 2265 22875 3.142 750 82.8 280 1279 1-Oct-99<br />

31-Dec-02<br />

Cosmos 2332 23853 3.242 750 82.9 286 1383 1-Oct-99<br />

31-Dec-02


ACCEPTED MANUSCRIPT<br />

Table 3<br />

Percentage by which density models are too high at sunspot maximum<br />

(percentage by which CD is too low)<br />

Altitude<br />

[km]<br />

J70 MOD JB2006<br />

150 4 7<br />

200 6 9<br />

250 8 11<br />

300 9 12<br />

350 10 13<br />

400 12 15<br />

500 12 15


ACCEPTED MANUSCRIPT<br />

Table 4<br />

Physical drag coefficients of Clementine<br />

Accommodation Physical drag coefficient<br />

Coefficient of Clementine<br />

1.00 2.281<br />

0.99 2.327<br />

0.98 2.364<br />

0.95 2.440<br />

0.90 2.531<br />

0.85 2.609


ACCEPTED MANUSCRIPT<br />

Table 5<br />

Drag and accommodation coefficients at sunspot maximum<br />

Altitude [km] Observed Drag Coefficient Corresponding <br />

200 2.08<br />

[Cosmos 1179, Cosmos 1427, Cosmos 1615]<br />

275 2.08<br />

[Cosmos 2265]<br />

2.14<br />

[Cosmos 2332]<br />

300 2.13<br />

[Cosmos 1179, Cosmos 1427, Cosmos 1615]<br />

2.16<br />

[Calsphere 3, Calsphere 4, Calsphere 5]<br />

400 2.24<br />

[Cosmos 1427, Cosmos 1615]<br />

2.24<br />

[Calsphere 3, Calsphere 4, Calsphere 5]<br />

480 2.485<br />

[SNOE]<br />

500 2.31<br />

[Calsphere 3, Calsphere 4, Calsphere 5]<br />

630 2.57<br />

[Clementine]<br />

1.00<br />

1.00<br />

0.993<br />

0.995<br />

0.990<br />

0.960<br />

0.960<br />

0.95<br />

0.915<br />

0.875


ACCEPTED MANUSCRIPT<br />

Table 6<br />

Accommodation coefficients at sunspot minimum<br />

Altitude<br />

[km]<br />

<br />

150 1.00<br />

175 1.00<br />

200 0.99<br />

225 0.98<br />

250 0.96<br />

275 0.94<br />

300 0.92<br />

325 0.89


Figure-1<br />

ACCEPTED MANUSCRIPT


Figure-2<br />

ACCEPTED MANUSCRIPT


Figure-3<br />

ACCEPTED MANUSCRIPT


Figure-4<br />

ACCEPTED MANUSCRIPT


Figure-5<br />

ACCEPTED MANUSCRIPT


Figure-6<br />

ACCEPTED MANUSCRIPT


Figure-7<br />

ACCEPTED MANUSCRIPT


Figure-8<br />

ACCEPTED MANUSCRIPT


Figure-9<br />

ACCEPTED MANUSCRIPT


Figure-10<br />

ACCEPTED MANUSCRIPT

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!