Table of Contents ECSS Model Page
Background Information Background Information
Meteoroid and debris models

Table of contents

1 Introduction

2 Environment models

3 Particle/wall interaction models

4 Implementation in SPENVIS

References

1 Introduction

The possibility of overcrowding space in the near Earth environment with meteoroids and orbiting man-made objects has been recognised as a real problem which already influcences space missions now and may well become a major aspect of future space flight in general. An example of direct influence is the shielding of all manned modules required in the International Space Station programme.

Proper design of the shielding concept requires the use of analysis tools which allow detailed risk assessment for given spacecraft, shielding design and mission. To eliminate approximations and uncertainties, such a prediction tool has to account fully for three-dimensional geometrical and directional effects. ESABASE, a multi-disciplinary computer-based tool for system engineering developed for ESA/ESTEC, includes an enhanced 3D numerical analysis tool for evaluation of the meteoroid and debris environments, impact probability and resulting damage effects. The meteoroid/debris tool implemented in SPENVIS does not include a full geometrical analysis, but does allow an estimation of the meteoroid and debris particle fluxes on randomly oriented surfaces in a user-defined orbit. The resulting fluxes can be fed into particle/wall interaction models to estimate the number of punctures over a given time interval.

2 Environment models

The environment models describe the particle flux relative to a moving spacecraft as a function of mass, velocity, direction, altitude and orbital inclination. For both meteoroids and space debris, the models include mass-density relations, directional collision velocity distributions and flux-mass/diameter relations. For meteoroids, flux corrections due to Earth are considered.

Meteoroid and space debris environments are presented separately because of their different properties and characteristics. The meteoroid environment encompasses only particles of natural origin whereas the space debris environment is man-made. The main differences in both classes of particle models are:

2.1 The meteoroid environment

The meteoroid environment encompasses only particles of natural origin. Meteoroids can be classified as sporadic (i.e. their number is independent of the time of the year) and streams, whose number increases by a factor 5 or more during certain periods of the year. The models consist of:

2.1.1 Mass-density models

As it is not a measured quantity, the density of meteoroids is open to serious uncertainty. Meteoroids are generally considered to be of cometary and asteroidal origin. The meteoroid density of interest here is that of particles which result from the break-up of cometary nuclei. The cometary meteoroid is described as a conglomerate of dust particles. values of density calculated from photographic and radar observations range from 0.16 g cm-3 to 4 g cm-3 [Cour-Palais, 1969].

The values for average mass density quoted in the literature vary widely, so that a value can only estimated. A value of 0.5 g cm-3 seems the most appropriate for a constant mass density (rho). A more appropriate mass-density relation is a decreasing and non-continuous function of mass given by:

  2.0 g cm-3 for m < 10-6 g
rho(m)  = 1.0 g cm-3 for 10-6 g <= m <= 10-2 g
  0.5 g cm-3 for m > 10-2 g

In SPENVIS, the user has the choice between this relation or a constant mass density.

2.1.2 The Grün interplanetary flux model

The flux-mass model for meteoroids at 1 AU of Grün et al. [1985] gives the total average meteoroid flux in terms of the integral flux, i.e. the number of particles per m2 per year of mass larger than or equal to a given mass m, impacting a randomly-oriented flat plate under a viewing angle of 2pi. Except for Earth shielding and gravitational effects, this flux is omnidirectional. This 'interplanetary' flux covers the range 10-18-1 g. It closely fits the fluxes from the HEOS 2 and PIONEER 8-9 measurements and is analytically described as:

F(m)  =  3.15576x107  [F1(m) + F2(m) + F3(m)] ,

where
F1(m) = (2.2x103 m0.306 + 15.0)-4.38

F2(m) = 1.3x10-9 (m + 1011 m2 + 1027 m4)-0.36

F3(m) = 1.3x10-16 (m + 106 m2)-0.85

Function F1(m) refers to large particles (m > 10-9 g), function F2(m) to intermediate sized particles (10-14 g < m < 10-9 g) and function F3(m) to small particles (m < 10-14 g). The Grün et al. (1985) model is represented in Fig. 1.

Grun flux-mass model
Figure 1. The Grün meteoroid flux-mass model

2.1.3 Flux corrections due to Earth

As the Grün model represents the total meteoroid flux at 1 AU from the Sun in the ecliptic plane, but in the absence of the Earth, two corrections need to be taken into account, one related to Earth shielding, which reduces the flux, and one to the gravitational effect of the Earth, which tends to increase the flux, in order to account for the presence of the Earth.
2.1.3.1 Earth shielding
The Earth shielding is accounted for via the geometrical multiplicative factor xi which depends on the spacecraft altitude. This has the effect of subtracting the flux within the solid angle subtended by the shielding body, the Earth shielding cone (see Fig. 2). If the target plate is assumed to be randomly oriented, an averaged xi factor is considered:

ximean = 0.5 (1 + cos theta)

where theta is defined by

sin theta = (RE + 100) / (RE + h)

RE being the mean Earth radius (6378 km) and h the altitude of the target above the Earth's surface in km (h => 100 km). The constant 100 corresponds to the atmosphere height above the Earth. The dependence of ximean on altitude is represented in Fig. 3.

Earth shielding effect
Figure 2. Earth shielding effect for meteoroid flux

Meteoroid geometrical Earth shielding factor
Figure 3. Geometrical shielding factor for meteoroid flux

2.1.3.2 Gravitational focussing
Due to the gravitational field of the Earth, meteoroid particles are attracted and therefore the flux increases compared to deep space. This effect is taken into account by the gravitational focussing factor GE defined as

GE = 1 + (RE + 100) / (RE + h)

for the Grün model. Figure 4 shows the graviational focussing factor as a function of altitude.

Gravitational focussing factor
Figure 4. Gravitational focussing factor for meteoroid flux

2.1.3.3 Combined correction
Consideration of the Earth shielding and gravitational focussing gives a particle flux-mass which depends on spacecraft altitude. This corrected flux Fc is expressed as

Fc(m,h) = F(m) GE(h) xi(h)

The combined correction factor GE(h) xi(h) is represented in Fig. 5.

Total meteoroid flux correction
Figure 5. Total meteoroid flux correction due to Earth

2.1.4 Meteoroid stream models

The Grün et al. [1985] meteoroid flux is defined by means of the cumulative sporadic plus streams flux averaged over one year. In fact, noticeable increases in the average hourly rate of meteor activity have been observed at regular intervals during the calendar year. These increases are caused by the Earth's passage through a stream of particles travelling in similar heliocentric orbits.

Two models of the annual meteoroid streams are available: the Cour-Palais [1969] method, and the Jenniskens [1994] method. In SPENVIS, no stream models are implemented.

2.1.4.1 The Cour-Palais model
The flux-mass model for stream meteoroids has been described by Cour-Palais [MBB-ERNO, 1988] as:

log Fst = -14.41 - log m - 4 log(vst/20) + log F

where Fst is the number of stream particles of mass m or larger in units of m-2 s-1, m the particle mass (g), Fst the geocentric velocity of the stream in km s-1, and F the integrated ratio of cumulative stream flux to the average cumulative sporadic flux.

The F function is a piecewise linear function of time that depends on five parameters:

Figure 6 shows a schematic representation of the F function. The stream parameters for major streams are listed in Table 1. The major streams are represented graphically in Fig. 7.

Meteoroid stream flux ratio
Figure 6. Meteoroid stream flux ratio

Table 1. Summary of the Cour-Palais characteristics of major stream events
Stream name Time of occurence
(day of year)
Duration
(days)
Increase rate
(deg)
Decrease rate
(deg)
Maximum flux ratio Velocity
(km s-1)
Quadrantids 2 2 82.87 7.13 8.0 42.0
Lyrids 109 3 23.03 49.64 0.85 48.0
eta Aquarids 121 7 36.25 42.27 2.2 64.0
omicron Cetids 134 9 90.0 90.0 2.0 37.0
Arietids 149 21 29.36 70.91 4.5 38.0
zeta Perseids 152 15 30.96 73.30 3.0 29.0
beta Taurids 175 11 26.57 74.05 2.0 31.0
delta Aquarids 207 10 36.87 79.38 1.5 40.0
Perseids 196 34 10.89 38.66 5.0 60.0
Orionids 288 10 11.31 39.80 1.2 66.0
Arietids Southern 274 58 3.27 85.03 2.0 28.0
Taurids Northern 299 27 1.53 88.09 0.4 29.0
Taurids Night 305 29 4.09 86.19 1.0 37.0
Taurids Southern 299 27 3.43 85.71 0.9 28.0
Leonids Southern 319 5 41.99 73.30 0.9 72.0
Bielids 316 4 11.31 78.69 0.4 16.0
Geminids 329 22 13.24 45.0 4.0 35.0
Ursids 354 4 51.34 38.66 2.5 37.0

Major meteoroid streams
Figure 7. The major Cour-Palais meteoroid streams

2.1.4.2 The Jenniskens model
The Jenniskens [1994] meteoroid stream model is based on data collected by a large number of observers over a 10 year period from observation sites in both the northern and southern hemispheres. The stream geometry and activity at shower maximum are defined by:
  1. the solar longitude lambdamax;
  2. the maximum zenithal hourly rate ZHRmax, which is the number of 'visible' meteors seen after various observer and location related corrections have been applied;
  3. the apparent radiant position in right ascension and declination of the radiant;
  4. the geocentric meteoroid speeds, defined as the final geocentric velocity vinfinity as the meteoroids reach the top of the atmosphere.
The right ascension (RA) and declination (DEC) of the radiant for an instantaneous values of the solar longitude are given by:

RA(lambda) = RA(lambda0) + DRA (lambda - lambda0)

and

Dec(lambda) = Dec(lambda0) + DDec (lambda - lambda0) ,

where DRA and DDec are the variations of RA and Dec per degree of solar longitude. The shower activity as a function of time around its maximum is described by:

ZHR = ZHRmax 10-B|lambda-lambdamax| ,

where B describes the slope of the activity profiles. Since most streams are found to have symmetrical profiles, a single values of B is sufficient. The Geminids are the exception: this stream needs a different value of B for the inward and outward slope. Six of the streams do not have a strong enough ZHR to produce a slope: here it is suggested to use a 'typical' value B = 0.2. Six other streams are best represented by the sum of two activity profiles, defined by a peak profile ZHRpmax and Bp and a background profile ZHRbmax with separate inward and outward slope values Bb+ and Bb-, respectively. This results in the following expression:

ZHR = ZHRpmax 10-Bp|lambda-lambdamax| + ZHRbmax (10-Bb-(lambdamax-lambda) + 10-Bb+(lambda-lambdamax))

The cumulative flux at solar longitude lambda can now be expressed as:

F(m,lambda) = F(m)max ZHR(lambda) / ZHRmax

with

F(m)max = k m-alpha ,

where F(m) is the cumulative flux (in m-2 s-1) of particles with mass greater than m (in kg), alpha is the cumulative mass distribution index, and k the cumulative mass distribution constant. The total particle flux Ftot is obtained by summation over all streams:

Ftot = Fsp + Sum Fst .

Since the Grün flux models all particles, including the streams, Ftot must be forced to equal the Grün flux when summed over a full year. Thus, when the stream model is used, the new sporadic flux becomes:

Fsp = FGrün - Sum Fst ,

where the sum is evaluated over one full year. Table 2 lists the parameters of the 50 Jenniskens streams.

Table 2. Parameters of the 50 Jenniskens streams
Stream name lambdamax RAmax DRA Decmax DDec ZHRpmax Bp+ Bp- ZHRbmax Bb+ Bb- alpha k vinfinity
Bootids 283.3 232 0.6 45 -0.3 110.0 2.50 2.50 20.0 0.37 0.45 0.92 8.4x10-17 43
gamma Velids 285.7 124 0.5 -47 -0.2 2.4 0.12 0.12 0.0 0.0 0.0 1.10 5.8x10-19 35
alpha Crucids 294.5 193 1.1 -63 -0.4 3.0 0.11 0.11 0.0 0.0 0.0 1.06 1.9x10-19 50
alpha Hydrusids 300.0 138 0.7 -13 -0.3 2.0 0.20 0.20 0.0 0.0 0.0 1.3 3.4x10-19 44
alpha Carinids 311.2 99 0.4 -54 0.0 2.3 0.16 0.16 0.0 0.0 0.0 0.92 1.3x10-17 25
delta Velids 318.0 127 0.5 -50 -0.3 1.3 0.20 0.20 0.0 0.0 0.0 1.10 3.1x10-19 35
alpha Centaurids 319.4 210 1.3 -58 -0.3 7.3 0.18 0.18 0.0 0.0 0.0 0.83 3.7x10-18 57
omicron Centaurids 323.4 176 0.9 -55 -0.4 2.2 0.15 0.15 0.0 0.0 0.0 1.03 1.9x10-19 51
theta Centaurids 334.0 220 1.1 -44 -0.4 4.5 0.20 0.20 0.0 0.0 0.0 0.95 4.4x10-19 60
delta Leonids 335.0 169 1.0 17 -0.3 1.1 0.049 0.049 0.0 0.0 0.0 1.10 1.9x10-18 23
Virginids 340.0 165 0.9 9.0 -0.2 1.5 0.20 0.20 0.0 0.0 0.0 1.10 1.5x10-18 26
gamma Normids 353.0 285 1.3 -56 -0.2 5.8 0.19 0.19 0.0 0.0 0.0 0.87 1.9x10-18 56
delta Pavonids 11.1 311 1.6 -63 -0.2 5.3 0.075 0.075 0.0 0.0 0.0 0.95 5.1x10-19 60
Lyrids 32.4 274 1.2 33 0.2 12.8 0.22 0.22 0.0 0.0 0.0 0.99 2.0x10-18 49
mu Virginids 40.0 230 0.5 -8 -0.3 2.2 0.045 0.045 0.0 0.0 0.0 1.10 1.1x10-18 30
eta Aquarids 46.5 340 0.9 -1 0.3 36.7 0.08 0.08 0.0 0.0 0.0 0.99 1.5x10-18 66
beta Corona Australids 56.0 284 1.3 -40 0.1 3.0 0.20 0.20 0.0 0.0 0.0 1.13 1.5x10-19 45
alpha Scorpiids 55.9 252 1.1 -27 -0.2 3.2 0.13 0.13 0.0 0.0 0.0 0.92 4.7x10-17 21
Da. Arietids 77.0 47 0.7 24 0.6 54.0 0.10 0.10 0.0 0.0 0.0 0.99 2.6x10-17 38
gamma Sagittarids 89.2 286 1.1 -25 0.1 2.4 0.037 0.037 0.0 0.0 0.0 1.06 1.9x10-18 29
tau Cetids 95.7 24 0.9 -12 0.4 3.6 0.18 0.18 0.0 0.0 0.0 0.92 3.7x10-19 6.6
theta Ophiuchids 98.0 292 1.1 -11 0.1 2.3 0.037 0.037 0.0 0.0 0.0 1.03 3.5x10-18 27
tau Aquarids 98.0 342 1.0 -12 0.4 7.1 0.24 0.24 0.0 0.0 0.0 0.92 8.9x10-19 63
nu Phoenicids 111.2 28 1.0 -40 0.5 5.0 0.25 0.25 0.0 0.0 0.0 1.10 2.6x10-19 48
omicron Cygnids 116.7 305 0.6 47 0.2 2.5 0.13 0.13 0.0 0.0 0.0 0.99 1.4x10-18 37
Capricornids 122.4 302 0.9 -10 0.3 2.2 0.041 0.041 0.0 0.0 0.0 0.69 8.3x10-17 25
tau Aquarids North 124.1 324 1.0 -8 0.2 1.0 0.063 0.063 0.0 0.0 0.0 1.19 3.6x10-20 42
Pisces Australes 124.4 339 1.0 -33 0.4 2.0 0.40 0.40 0.9 0.03 0.10 1.16 1.5x10-19 42
delta Aquarids Southern 125.6 340 0.8 -17 0.2 11.4 0.091 0.091 0.0 0.0 0.0 1.19 3.6x10-19 43
iota Aquarids Southern 131.7 335 1.0 -15 0.3 1.5 0.07 0.07 0.0 0.0 0.0 1.19 1.2x10-19 36
Perseids 140.2 47 1.3 58 0.1 70.0 0.35 0.35 23.0 0.05 0.092 0.92 1.2x10-17 61
kappa Cygnids 146.7 290 0.6 52 0.3 2.3 0.069 0.069 0.0 0.0 0.0 0.79 3.0x10-17 27
pi Eridanids 153.0 51 0.8 -16 0.3 40.0 0.20 0.20 0.0 0.0 0.0 1.03 1.7x10-18 59
gamma Doradids 155.7 60 0.5 -50 0.0 4.8 0.18 0.18 0.0 0.0 0.0 1.03 1.1x10-18 41
Aurigids 158.2 73 1.0 43 0.2 9.0 0.19 0.19 0.0 0.0 0.0 0.99 2.9x10-19 69
kappa Aquarids 177.2 339 0.9 -5 0.4 2.7 0.11 0.11 0.0 0.0 0.0 1.03 1.9x10-17 19
epsilon Geminids 206.7 104 0.7 28 0.1 2.9 0.082 0.082 0.0 0.0 0.0 1.10 2.1x10-20 71
Orionids 208.6 96 0.7 16 0.1 25 0.12 0.12 0.0 0.0 0.0 1.13 1.6x10-19 67
Leo Minorids 209.7 161 1.0 38 -0.4 1.9 0.14 0.14 0.0 0.0 0.0 0.99 1.1x10-19 61
Taurids 223.6 50 0.3 18 0.1 7.3 0.026 0.026 0.0 0.0 0.0 0.83 4.3x10-17 30
delta Eridanids 229.0 54 0.9 -2 0.2 0.9 0.20 0.20 0.0 0.0 0.0 1.03 7.5x10-19 31
zeta Puppids 232.2 117 0.7 -42 -0.2 3.2 0.13 0.13 0.0 0.0 0.0 1.22 9.5x10-20 41
Leonids 235.1 154 1.0 22 0.4 19.0 0.55 0.55 4.0 0.025 0.15 1.22 3.4x10-20 71
Puppids/Vel 252.0 128 0.8 -42 -0.4 4.5 0.034 0.034 0.0 0.0 0.0 1.06 8.2x10-19 40
Phoenicids 252.4 19 0.8 -58 0.4 2.8 0.30 0.30 0.0 0.0 0.0 1.03 2.5x10-17 18
Monocerotids 260.9 100 1.0 14 -0.1 2.0 0.25 0.25 0.0 0.0 0.0 1.25 3.3x10-20 43
Geminids 262.1 113 1.0 32 0.1 74.0 0.59 0.81 18.0 0.09 0.31 0.95 7.8x10-17 36
sigma Hydrusids 265.5 133 0.9 0 -0.3 2.5 0.1 0.1 0.0 0.0 0.0 1.10 4.7x10-20 59
Ursids 271.0 224 -0.2 78 -0.3 10.0 0.90 0.90 2.0 0.08 0.2 1.22 8.1x10-19 35

2.1.5 Further directional effects

McBride and McDonnell [1996] have shown that actually very little is known about how the Grün flux should be modified to include an apex enhancement, to sort out the beta meteoroids, and to include interstellar dust.
2.1.5.1 Separation of alpha and beta source
McBride and McDonnell [1996] suggest to separate the Grün flux into an alpha population and a beta population which has a crossover at 10-11 g. The beta population has the direction from the Sun and is of the small particle size. The separation into alpha flux and beta flux then is given by:

Fbeta(m) = FGrün(m) - Falpha(m) ,

with

Falpha(m) = FGrün(m) FH(m) / [FGrün(m) + FH(m)]

and

log FH(m) = -0.146 log m - 6.427 .

With the above equations it is possible to calculate for each mass the corresponding cumulative fluxes Falpha and Fbeta from the Grün flux FGrün. Figure 8 shows the separation of the Grün flux into alpha and beta components.

The velocity of the beta particles is size dependent:

v(m) = v0 (1011 m)-gamma ,

with v0 = 20 km s-1 and gamma = 0.18.

alpha And beta components of the Grun flux
Figure 8. alpha And beta components of the Grün isotropic meteoroid flux

2.1.5.2 Apex enhancement of the alpha source
McBride and McDonnell [1996] use a minimum to maximum antapex to apex flux ratio RF (which is in fact unknown) to define a modulation of the flux and of the velocity about the apex direction. The angular deviation from the apex direction is denoted by t, and it is assumed that a parameter delta describes a slight deviation from the measured peak value which was observed to be about 10° off the apex direction. Thus, the modulation of the alpha flux and velocity can be defined as:

Falpha(t) = F0alpha [1 + DeltaF cos(t+delta)]

valpha(t) = v0alpha [1 + Deltav cos(t+delta)] ,

where

DeltaF = (RF - 1) / (RF + 1) ,

Deltav = (vA - vAA) / (vA + vAA) ,

v0alpha = (vA + vAA) / 2 ,

and the subscripts A and AA denote apex and antapex, respectively.

From the AMOR meteor data there are some guesses for the maximum to minimum detectrion ratio from which one may try to obtain values for RF and vAA. Although RF could be anywhere in the range 1 to 5 and vA and vAA are not known either, it is recommended to use the following values for a first guess: vA = 17.7 km s-1, vAA = 8.3 km s-1, RF = 2, resulting in the values DeltaF = 0.33 and Deltav = 0.36.

2.1.5.3 Interstellar dust
McBride and McDonnell [1996] have identified two components of interstellar dust.

The first source concerns measurements on Ulysses and Galileo, which detected at about 5 AU particles of 3x10-16 g with heliocentric velocities of 26 km s-1, ecliptic longitude 252° and latitude 2.5°. The total particle flux at 1 AU is estimated to be 5x10-4 m-2 s-1 and the heliocentric velocity 47 km s-1.

The second source [Taylor et al., 1996] stems from AMOR meteor data which indicate at least two sources, defined by:

The mass of these interstellar meteoroids is estimated to lie between 15 and 40 ??. No flux is known for these contributions.

2.1.6 Velocity distributions

Three velocity density distributions, relative to the Earth, are available in the literature: the first one has been defined by Cour-Palais [1969], the second one is part of the NASA90 model Anderson [1991], and the third is due to Taylor [1995].
2.1.6.1 The Cour-Palais distribution
In the Cour-Palais model, the velocity v ranges from 11 to 72 km s-1 with a mean value of 20 km s-1. The Cour-Palais velocity density distribution, originally defined in histogram form, can be smoothed by a function g (in km s-1) with a mean value 20 km s-1 and normalised over the velocity interval from 0 to infinity:

g(v)  =
4/81 (v - 11) exp[-2/9 (v-11)] for v > 11
0 for v <= 11

The Cour-Palais velocity density distribution is represented in Fig. 9.

Cour-Palais meteoroid velocity density
Figure 9. Cour-Palais meteoroid velocity density

2.1.6.2 The NASA90 distribution
In the NASA90 model, the velocity v ranges from 11.1 to 72.2 km s-1 with a mean value of 17 km s-1. The NASA90 velocity density distribution is defined as

g(v)  =
0.112 for 11.1 <= v < 16.3
3.328x105 v-5.34 for 16.3 <= v < 55
1.695x10-4 for 55 <= v <= 72.2

The NASA90 velocity density distribution is represented in Fig. 10.

NASA90 meteoroid velocity density
Figure 10. NASA90 meteoroid velocity density

2.1.6.3 The Taylor distribution
The velocity distributions of meteoroids at 1 AU (i.e. as viewed from a massless Earth) have generally been derived from ground based observations of photographic meteors, which are corrected for the effect of the Earth gravity. McBride and McDonnell [1996] have compared distributions by Erickson, Sekania and Southworth and by Taylor. They concluded that the most statistically reliable published data set comes from the Harvard Radio Meteor Project (HRMP) in which about 20,000 meteor observations were evaluated. Taylor [1995] re-evaluated and corrected the original measurement data and derived the normalised distribution shown in Fig. 11.

Taylor meteoroid velocity density
Figure 11. Taylor meteoroid velocity density

2.2 The space debris environment

In contrast to the meteoroid environment, the space debris environment is man made. It consists of inactive satellites or spacecraft, non-operational payloads and fragments from separation or deployment procedures, accidental rocket explosions or anti-satellite tests. The orbital debris environment is more hazardous to space flight than the meteoroid environment below 2,000 km.

Kessler established that the total mass of space debris below 2,000 km is far larger than the meteoroid material. Most of the debris particles are in high inclination orbits with an average speed of 10 km s-1. Figure 12 shows the predicted mass distribution of space debris resulting from an upperstage explosion. Figure 13 shows the average flux as a function of mass between 700 and 1,200 km for different years. It is estimated that the amount of mass with diameter 1 cm and smaller will increase at a rate of 2 to 10% per year. This effect is accounted for in the debris models.

Predicted mass distribution for an upperstage explosion
Figure 12. Predicted mass distribution of space debris resulting from an upperstage explosion

Average debris flux between 700 and 1,200 km
Figure 13. Average debris flux between 700 and 1,200 km for different years

The description of the space debris environment is useful for low Earth orbit, because a large portion of the debris population has small sizes. For geostationary orbits, which are relevant for telecommunication satellites, a different approach is used. A sufficiently large portion of objects are tracked individually (e.g. by NORAD). From this data base, a collision probability is derived using statistical methods [McCormick, 1986; Hechler, 1985]. Nevertheless, small sized objects, which cannot be tracked, lead to uncertainties and increases in the collison probability.

Due to their origin, space debris particles have the following properties:

Most debris is in circular orbits.

Three space debris environment models are described below:

2.2.1 The NASA90 model

The NASA90 debris model is an analytical formulation of the flux vs. particle diameter. The orbit altitude, epoch and inclination are used in the formulation. The main input parameter is the particle diameter. The increase of space debris is accounted for by user defined growth rates. The flux formulation computes the impact flux on a random tumbling plate.
2.2.1.1 The NASA90 flux model
The flux F, which is the cumulative number of impacts on a spacecraft in circular orbit per m2 and per year on a randomly tumbling surface, is defined as a function of the minimum debris diameter d (in cm), the target orbit altitude h (in km, with h<=2,000 km), the target orbit inclination i (in degrees), the mission date t (in years), the 13 month average solar radio flux S (in 104 Jy) measured in the year prior to the mission, the annual growth rate p of mass in orbit (default 0.05), and the growth rate q of fragment mass (default: 0.02 before 2011, 0.04 after 2011):

F(d,h,i,t,S) = H(d) Phi(h,S) Psi(i) [F1(d) g1(t,p) + F2(d) g2(t,p)] ,

with

H(d) = 100.5 exp[-(log d - 0.78)2 / 0.406]

Phi(h,S) = Phi1(h,S) / [1 + Phi1(h,S)]

Phi1(h,S) = 10h/200 - S/140 - 1.5

F1(d) = 1.22x10-5 d-2.5

F2(d) = 8.1x1010 (d + 700)-6

g1(t,q) = (1 + q)t-1988

g2(t,p) = 1 + p (t-1988)


The function Psi(i) defines the relationship between the flux on a spacecraft in an orbit of inclination i and the flux incident on a spacecraft in the current population's average inclination of about 60°. Psi(i) is tabulated in Table 3; for intermediate values of i, a linear interpolation is performed.

Table 3. NASA90 Inclination dependent function Psi
i (deg) Psi(i)
<=28.50.91
300.92
400.96
501.02
601.09
701.26
801.71
901.37
1001.78
=>1201.18

Figure 14 shows the NASA90 flux model for altitude 400 km and inclination 51.6°.

NASA90 Flux
Figure 14. NASA90 Flux for altitude 400 km and inclination 51.6°

2.2.1.2 The NASA90 velocity distribution
The collision velocity distribution g(v,i,h), i.e. the number of impacts with velocity between v and v+dv, relative to a moving spacecraft with orbital inclination i and altitude h, is given by:

g(v,i,h) = v [2v0(i,h) - v{ g1(i) exp{-[(v - 2.5v0(i,h)) / (g2(i) v0(i,h))]2}
+ g3(i) exp{-[(v - g4(i) v0(i,h)) / (g5(i) v0(i,h))]2}
+ g6(i) v [4v0(i,h) - v] ,

where the functions gj(i) and v0(i,h) are piecewise polynomials defined as:

g1(i)  =  
18.7    i < 60°
18.7 + 0.0298 (i - 60)3    60° <= i < 80°
250    i => 80°
 
g2(i)  =  
0.5    i < 60°
0.5 - 0.01 (i - 60)    60° <= i < 80°
0.3    i => 80°
 
g3(i)  =  
0.3 + 0.0008 (i - 50)2    i < 50°
0.3 - 0.01 (i - 50)    50° <= i < 80°
0    i => 80°
 
g4(i)  =   1.3 - 0.01 (i - 30)
 
g5(i)  =   0.55 + 0.005 (i - 30)
 
g6(i)  =  
0.0125 [1 - 0.0000757 (i - 60)2]    i < 100°
[0.0125 + 0.00125 (i - 100)] [1 - 0.0000757 (i - 60)2]    i => 100°
 
v0(i,h)  =  
v0(h) [7.25 + 0.015 (i - 30)] / 7.7    i < 60°
v0(h)    i => 60°

and v0(h) is the velocity at target orbit altitude h:

v0(h) = 631.7 (6378 + h)-0.5

Figure 15 shows the NASA90 velocity distribution for altitude 400 km and inclination 51.6°.

NASA90 Velocity distribution
Figure 15. NASA90 Velocity distribution for altitude 400 km and inclination 51.6°

2.2.1.3 The NASA90 mass density model
For the NASA90 model the particle mass density rho can be either set to a constant with default value 2.8 g cm-3 or the following relation can be used:

rho(d) =  
2.8 d-0.74    d => 0.62
4.0    d < 0.62

where d is the particle diameter (in cm).

2.2.3 The NASA96 model

The Johsnon Space Centre [Kessler et al., 1996] has published a new engineering model, ORDEM96, which constitutes an update to the NASA90 model. It will be referenced as the NASA96 model. This model is based on various empirical considerations and basically defines orbits in six different inclination domains. For each orbit, a numerical collision analysis needs to be made, which yields a spatial debris density around the target orbiter together with directional information, which leads to debris fluxes on the specific target orbit. Unlike the NASA90 model, the NASA96 model does not directly define fluxes but defines numbers of orbits of various contribution types which first need to be numerically converted to fluxes.
2.2.3.1 General description
In the NASA90 model, the debris fluxes and the velocity distributions are defined by empirical functions which depend on debris diameter, altitude, epoch, and inclination. In the NASA96 model it is no longer possible to express the flux and velocity distributions explicitly as a function of orbital elements.

Due to more refined measurement data, primarily produced by Haystack radar and LDEF, it became necessary for the NASA96 model to define debris orbit numbers of circular and elliptical type with given particle size, altitude, and perigee altitude distributions, for six inclination ranges. With this orbit information it is then possible to conduct a numerical collison analysis which yields the flux on a given target orbit by calculating the spatial debris densitites along the target orbit.

In addition to circular orbits, eccentric debris orbits are now included. Thus, impact directions with non vanishing elevations are obtained. The circular orbit family represents debris orbits with eccentricity e < 0.2, while the eccentric orbit family represents debris orbits with eccentricity e > 0.2. The apogee of the elliptic orbits is fixed at 20,000 km, but the perigee altitudes vary.

Six source components are used, in decreasing size:

These size ranges only indicate the dominating contributions, i.e. the various debris sources also contribute outside the indicated ranges.

The functional forms used to represent the number of particles consist of 12 sets of equations, which correspond to the six inclination bands with each a set of circular and elliptical orbits, and with the following functional dependencies of the particle sizes:

2.2.3.2 Definition of the debris orbit distributions
Circular and eccentric orbits are defined in the following six inclination ranges (the numbers in brackets are the central values): 0° <= i < 19° (7°), 19° <= i < 36° (28°), 36° <= i < 61° (51°), 61° <= i < 73° (65°), 73° <= i < 91° (82°), 91° <= i < 180° (98°). The eccentric orbits all have the same apogee altitude of 20,000 km.

The general forms of the equations for the number of debris objects within an altitude or perigee altitude bin of 1 km width, for a particular inclination range, are:

Nc = 0.1 g(t) [ phii(h,s) Fi(d) + phif(h,s) Ff(d) + phis(h,d,s) Fs(d) + phip(h) Fp(d) + phim(h,s) Fm(d) ]

for circular orbits, and

Ne = g(t) [ phii(q) Fi(d) + phif(q) Ff(d) + phis(q) Fs(d) + phip(q) Fp(d) + phim(q) Fm(d) ]

for elliptical orbits, where h is the circular orbit altitude and q the eccentric orbit perigee altitude, s is the solar activity F10.7 in the year prior to the mission, and the indices refer to the size ranges defined above. The growth factor g(t) is defined as:

g(t) = 1 + g0 (t - 1995) ,

with g0 between 0.04 and 0.08 and t the mission time in calendar year. Some of the phii and phif functions further depend on the production rate of new debris to the historical rate, which takes values between 0.2 and 1.0. The functions phi are sums of rational functions of exponential functions with exponents that are linear functions of h or q. The factors F are rational functions of powers of d.
2.2.3.3 Numerical collision analysis
Kessler et al. [1996] suggest to determine the flux on a given target orbit as a function of debris diameter and the azimut and elevation in 1 km bins of altitude or perigee altitude by performing a numerical collison analysis based on the calculation of spatial densities.
2.2.3.4 Limitations of the NASA96 model
The NASA96 model is essentially limited to low Earth and nearly circular target orbits. It does not, for instance, predict a collison probability along geostationary orbits or on geostationary transfer orbits, except for the near perigee region.

The collision analysis implies that fluxes are averaged over one orbital revolution of the spacecraft, so that no flux variations along the orbital path are seen. In addition, one must be aware that the collison analysis uses spatial densities which are functions of the distance and latitude only, because the flux is averaged over the argument of perigee and over the right ascension of the ascending node.

Only two eccentricity families are considered in the NASA96 model. Debris orbits of eccentricity lower than 0.2 are represented by circular orbit part, while debris orbits with eccentricities above 0.2 are represented by elliptical orbits with fixed apogee height of 20,000 km. Thus, the circular orbit part does not give any contributions on the top and bottom surface impacts by orbits in the eccentricity range 0.0-0.2. For higher eccentricities it is possible to obtain impact probabillities on these surfaces, but one must be aware that this does not represent the full contribution.

Molniya type heliosynchronous orbis, with their considerable contributions of delta upper stae explosion event debris, are not yet included in the NASA96 model. However, for target orbiters with inclinations lower than 63° this is of no concern.

2.2.4 The MASTER model

Initiated via an ESA contract [Sdunnus, 1995], a numerical data base MASTER was established. Fragmentation models have been applied to all known (127 as of 1996) break-up events. This resulted in a total of about 1011 objects above 100x10-6 m in size, which was reduced by sampling techniques to about 250,000 objects. Together with the trackable population, which is represented by the two line element data set, the orbits of these objects are propagated to a reference epoch. The resulting spatial density is stored in bins at the reference epoch. Sophisticated data compression techniques are employed to store the large amount of data.

The MASTER lower mass threshold is 0.1 mm. The altitude ranges from low Earth orbit (186 km) to geostationary altitude (35,800 km). The MASTER data base is available on CD-ROM and is regularly updated.

MASTER 2005
This version is recommended by ECSS standards.
MASTER 2009
A new enhanced MASTER model has been developed and was released in 2009. Simulation of the space debris environment in the new ESA's MASTER Model (Meteoroid and Space Debris Environment Reference) is performed through modelling of the physical generation processes for the different sources. The following sources of debris are considered: launch and mission-related objects, explosion and collision fragments, solid rocket motor slag and dust, NaK droplets, surface degradation products, ejecta, and meteoroids. MASTER can deliver flux and spatial density analysis for all epochs between 1957 and 2060. For all epochs, the lower size threshold is one micron. The analysis of the future debris environment (after 2009) is possible based on three different future scenarios (business as usual, intermediate mitigation, full mitigation).

3 Particle/wall interaction models

For the micro-particle environment risk analysis, the interaction of the particles with the spacecraft structures is as important as the evaluation of the particle fluxes. For this purpose, so-called damage equations, describing the physical phenomena of a particle impacting a structure at very high velocities, are used. The typical impact velocity for space debris is 8 to 10 km s-1, for meteoroids about 20 km s-1.

Due to the very complex physics of the impact phenomena, e.g. melting or shattering of the impactor, the equations are derived from experiments using empirical formulae. Lately, hydro-codes have been used to simulate hypervelocity impacts. However, due to the huge requirements in computation power and the delicate operation of these codes, they have not yet become standard for this type of phenomena.

The damage equations are treated in two separate groups:

Four classes of damage equations are described below:

Use of the damage equations needs information on the target design. Two designs are considered:

3.1 The damage equations

In order to provide the necessary flexibility in the usage of currently available and possible future damage equation formulations, the five classes of damage equations have been formulated in a parametric form, allowing the user to adapt the equation to his needs. The parameters for the standard equations are listed with each equation class.

The ballistic limit equations are used to compute the critical particle diameter, and the damage size equations to compute the crater or hole diameter. The symbols used in the equations are listed in Table 4.

Table 4. General symbols for the damage equations
Symbol Units Description
tt, tb, ts cm Thickness of target, back-up wall, shield
K   Characteristic factor
dp cm Particle (impactor) diameter
rhot, rhop, rhos, rhob g cm-3 Density of target, particle, shield, back-up wall
v km s-1 Impact velocity
alpha   Impact angle
S cm Spacing between shielding and back-up wall
D cm Crater or hole diameter

3.1.1 The single wall ballistic limit equation

The parametric formulation of the single wall ballistic limit equation is:

dp,lim = { tt / [Kf K1 rhopbeta vgamma cosxialpha rhotkappa] }1/lambda

The Kf factor allow to specify what type of damage is considered a failure for the thick plate equation and the glass target equations. In the other equations, it is not used. The K1 factor includes other parameters particular to each equation (e.g. target yield strength sigmat). The values of the parameters for the standard equations are given in Table 5.

Table 5. Standard parameter values for the single wall ballistic limit equation
Equation Kf (1) K1 (2) lambda beta gamma xi kappa
Thick plate 1.8-3 0.2-0.33 1.056 0.519 2/3 2/3 0
Thin plate 1.0 0.26-0.64 1.056 0.519 0.875 0.875 0
MLI (3) 1.0 0.37 1.056 0.519 0.875 0.875 0
Pailer-Grün 1.0 0.77 1.212 0.737 0.875 0.875 -0.5
McDonnell & Sullivan 1.0 0.756 [sigmaAl / sigmat]0.134 1.056 0.476 0.806 0.806 -0.476
Gardner 1.0 0.608 sigmat-0.093 1.059 0.686 0.976 0.976 -0.343
Gardner, McDonnell, Collier 1.0 0.85 sigmat-0.153 1.056 0.763 0.763 0.763 -0.382
Frost 1.0 0.43 1.056 0.519 0.875 0.875 0
Naumann, Jex, Johnson 1.0 0.65 1.056 0.5 0.875 0.875 -0.5
Naumann 1.0 0.326 1.056 0.499 2/3 2/3 0
McHugh & Richardson thick glass target 1.85-7 0.64 1.2 0 2/3 2/3 0.5
Cour-Palais thick glass target 1.85-7 0.53 1.06 0.5 2/3 2/3 0
(1) Failure factors Kf:
    - Thick plate: Kf => 3 Crater without spall
      2.2 <= Kf < 3 Spallation of the plate
      1.8 <= Kf < 2.2 Spall breaks away
      Kf < 1.8 Perforation of the plate
    - Thick glass targets: Kf => 7 Crater generation without spall
      1.85 <= Kf < 7 Spallation of the plate
      Kf < 1.85 Perforation of the plate
(2) K1 factors:
    - Thick plate: Aluminum alloys: K1 = 0.33
      Stainless steel: K1 = 0.2
    - Thin plate: Aluminum alloys: K1 = 0.43-0.454
      Stainless steel: K1 = 0.255  AISI 304, AISI 306
        K1 = 0.302  17-4 PH annealed
      Magnesium lithium: K1 = 0.637
      Columbium alloys: K1 = 0.271
    - McDonnell & Sullivan: Reference sigmat values given below
    - Gardner: Input sigmat in Pa for this equation
(3) The single wall ballistic limit equation for MLI assesses the failure of the thermal blanket and was derived by tests and hydro-code simulations.
Reference values for the 0.2 yield strength sigma used in the McDonnell & Sullivan and Gardner equations
Material xi (1) MPa
Aluminium pure 10 70
Aluminium alloys (superior) 30-65 200-450
Silver 22 150
Gold 17.5 120
Beryllium copper 120 830
Copper 32 220
Stainless steel 110 760
Titanium 140 980
(1) xi = 1,000 lb in-2 = 6.895 MPa

3.1.2 The multiple wall ballistic limit equation

The parametric formulation of the multiple wall ballistic limit equation is:

dp,lim = { [tb + K2 tsmu rhosv2] / [K1 rhopbeta vgamma cosxialpha rhobkappa Sdelta rhosv1] }1/lambda

Three velocity regions are defined, delimited by the two limit velocities vlim1 and vlim2. The governing parameters mostly have different values for velocities below vlim1 and above vlim2. For velocities between vlim1 and vlim2, a linear interpolation is performed. The limit velocities may vary with impact angle:

vlim1 = vlim1,0 cosphi1 alpha
vlim2 = vlim2,0 cosphi2 alpha

When the normal velocity component is used (which is generally the case), i.e. gamma = xi, the cosine exponents in the equations above are phi1 = phi2 = -1.

The parameter values for typical multiple wall equations are given in Tables 6 and 7 for double walls and multiple walls, respectively. For the Cour-Palais, MLI and Maiden-McMillan equations, only one velocity domain is used.

Table 6. Standard parameter values for the double wall ballistic limit equation
Equation K1 (1) K2 (2) lambda beta gamma kappa delta xi v1 / v2 mu
Cour-Palais 0.044 (sigmay,ref/sigmay,t)0.5 0 1 0.5 1 0.167 -0.5 1 0 / 0 0
MLI (3) 0.034 (sigmay,ref/sigmay,t)0.5 0 1 0.5 1 0.167 -0.5 1 0 / 0 0
Maiden-McMillan (1) Kf pi/6 (sigmay,ref/sigmay,t)0.5 0 3 1 1 0 -2 1 0 / 0 0
ESA   (2)  v < 3
                v > 9.5
0.255 +/- 0.637
pi/6 (sigmay,ref/sigmay,t)0.5
1
0
1.056
3
0.519
1
0.875
1
0
0
0
0
0.875
1
0 / 0
0 / 0
1
0
(1) Failure factors Kf for the Maiden-McMillan equation:
    Kf => 41.5 No damage
    8.2 <= Kf < 41.5 Incipient yield zone
    7.1 <= Kf < 8.2 Fracture zone
    Kf < 7.1 Penetration zone
(2) ESA Equation:
    - The Boeing-ESA equation has the same form as the ESA equation, but with vlim1 = 1.4 km s-1 and vlim2 = 7.83 km s-1.
    - The reference yield strength sigmay,ref = 70,000 lb in-2 = 482.8 MPa.
(3) MLI Equation:
    The multiple wall ballistic limit equation for MLI asses the debris/meteoroid protection of the thermal blanket, and was derived by tests and hycro-code simulations using the Cour-Palais equation as starting point.

Table 7. Standard parameter values for the multiple wall ballistic limit equation
Equation K1 K2 lambda beta gamma kappa delta xi v1 / v2 mu
ESA         v < 3
Triple       v > 7
0.312 (tau1* / tau)0.5
0.107 (tau1* / tau)0.5
1.667 K1
0
1.056
1.5
0.5
0.5
2/3
1
0
0
0
-0.5
5/3
1
0 / 0
0.167 / 0
1
0
NASA      v < 3
ISS           v > 7
0.6 (sigmaw / 40)-0.5
[3.918 (sigmaw / 70)1/3]-1.5
(sigmaw / 40)-0.5
0
1.056
1.5
0.5
0.5
2/3
1
0
0
0
-0.5
5/3
1
0 / 0
0.167 / 0
1
0
NASA      v < 3
Shock       v > 6
0.3 (tau1* / tau)0.5
22.545 (tau1* / tau)0.5
1.233 K1
0
1.056
3
0.5
1
2/3
1
0
-1
0
-2
5/3
1
0 / 1
0 / 0
1
0
NASA      v < 3
Bumper    v > 6
0.4 (tau1* / tau)0.5
18.224 (tau1* / tau)0.5
0.925 K1
0
1.056
3
0.5
1
2/3
1
0
-1
0
-2
5/3
1
0 / 1
0 / 0
1
0
tau1* Is the yield stress of a reference material (higher quality aluminium): tau1* = 40,000 lb in-2 = 276 MPa.
sigmaw = 47 xi for the reference equation used for system tests (to be input in xi).

3.1.3 The crater size equation

The parametric formulation of the crater size equation is very similar to the single wall ballistic limit equation:

D = K1 Kc dplambda rhopbeta vgamma cosxialpha rhotkappa

target cratering The crater factor Kc is the ratio of the crater radius D/2 to the crater depth p. The Kc factor can be classed in ductile and brittle targets. For ductile targets, the crater is more or less spherical and Kc is about 1. For brittle targets, an interior crater with diameter Dh may form, the outer crater being much larger. For brittle targets, Kc may be as high as 10.

Strictly speaking, the crater size equation is only valid when no failure occurs. The crater size equation assumes a semi-infinite target and should only be used for cases where the wall thickness is much larger than the particle diameter.

The values of the parameters for the standard equations are given in Table 8.

Table 8. Standard parameter values for the crater size equation
Equation Kc K1 lambda beta gamma xi kappa
Ductile targets
Thick plate (1) 2 0.4 +/- 0.66 1.056 0.519 2/3 2/3 0
Shanbing et al. n/d (2) 0.54 sigmay,t-1/3 1 2/3 2/3 2/3 -1/3
Sorensen n/d (2) 0.622 tau1-0.282 1 0.167 0.564 0.564 0.115
Christiansen (3) for rhop / rhot < 1.5 n/d (2) 10.5 Ht-0.25 cs2/3 1.056 0.5 2/3 2/3 -0.5
Christiansen (3) for rhop / rhot > 1.5 n/d (2) 10.5 Ht-0.25 cs2/3 1.056 2/3 2/3 2/3 -2/3
Brittle targets
Gault n/d (2) 1.08 1.071 0.524 0.714 0.714 -0.5
Fechtig n/d (2) 6.0 1.13 0.71 0.755 0.755 -0.5
McHugh & Richardson n/d (2) 1.28 1.2 0 2/3 2/3 0.5
Cour-Palais n/d (2) 1.06 1.06 0.5 2/3 2/3 0
(1) K1 factors:
    - Thick plate: Aluminium alloys: K1 = 0.66
      Stainless steel: K1 = 0.4
(2) n/d Means not defined in the equation reference. Default values of 1 and 10 are used for ductile and brittle targets, respectively.
(3) Christiansen equations:
    - Ht Is the target Brinell hardness. A typical value is 90.
    - cc Is the velocity of sound in the target material. For steel, cs = 5.85 km s-1.

3.1.4 The clear hole equation

The parametric formulation of the clear hole equation is:

D = [K0 (ts / dp)lambda rhopbeta vgamma cosxialpha rhosnu + A] dp

The clear hole equation is only valid for a full perforation, i.e. mainly for thin foils (typically bumper shields or similar). The limit of validity is given by the relation ts / dp < 10.

The values of the parameters for the standard equations are given in Table 9.

Table 9. Standard parameter values for the clear hole equation
Equation K0 lambda beta gamma xi nu A
Maiden 0.88 2/3 0 1 1 0 0.9
Nysmith-Denardo 0.88 0.45 0.5 0.5 0.5 0 0
Sawle 0.209 2/3 0.2 0.2 0.2 -0.2 1
Fechtig 5.24x10-5 0 1/3 2/3 2/3 0 0

3.1.5 Secondary ejecta

The process of secondary ejecta produced by hypervelocity impacts is relevant to more complex spacecraft structures. Examples are ejecta from solar arrays on the main satellite structure, or from baffles and other structures.
3.1.5.1 Normal impacts
Material ejection under hupervelocity impact is divided in three processes corresponding to different physical and mechanical phenomena: jetting phase, debris cone formation and spallation. In general, no spalls are observed on ductile targets. Figure 16 shows a schematic representation of ejection processes under normal impact.

Ejection processes under normal impact
Figure 16. Schematic representation of ejection processes under normal impact

The jetting phase
At the onset of penetration of the projectile into the target, both target and projectile undergo partial or complete melting and vaporisation. A certain amount of material is ejected from the impact interface. The physical state of the ejected material is mainly liquid and the ejection angle is approximately 20° measured from the target surface. The ratio of jetted mass to toal ejected mass is very small, less than 1%.
The debris cone
Later in the crater formation, the target material is finely commuted in fine solid fragments by compression or tensile failure. These fragments are ejected in a thin debris cone. The physical state of the ejected material is mainly solid. The ejection angle is between 60° and 80° measured from the target surface and depends on the target characteristics. The ratio of ejected mass to total ejected mass is estimated between 50% and 70%. The ejection velocity from a few m s-1 to a few km s-1 is inversely proportional to fragment size. The minimum size depends mainly on target characteristics and should be sub-micron sized. The maximum size can be evaluated by empirical relations. The size distribution is inversely proportional to the square of the fragment size.
The spallation phase
Near the free surface, rarefaction waves produce tensile stress. In brittle material, tensile failure leads to the formation of spall fragments that are ejected. The physical state of the ejected material is mainl solid. The ejection direction is normal to the surface. The ejection velocity is less than 1 km s-1 and is 10 to 100 times less than the impact velocity. The fragment size is large, about 10 times the size of debris cone fragments. These fragments have plate shapes whose dimensions are difficult to evaluate because large plate fragments are likely to fragment themselves into smaller particles. The ratio of spalled mass to total mass is estimated between 30% and 50%. In general, no spallation is observed on ductile targets.
3.1.5.2 Oblique impacts
Two cases have to be considered separately: The impact angles are measured from the target surface. Figure 17 shows a schematic representation of ejection processes under oblique impact.

Ejection processes under oblique impact
Figure 17. Schematic representation of ejection processes under oblique impact

Incidence impact angle > 30°
The phenomena involved in ejecta formation are similar to those observed for normal impacts: jetting phase, debris cone formation and spallation phase. The main differences are:
Incidence impact angle < 30°
For grazing incidence angles, the ricochet phenomenon appears and changes completely the physical and mechanical behaviour of both target and projectile. The projectile is probably not entirely consumed during impact and produces large high velocity solid fragments. The target is less compressed but is torn in the direction of impact. Experiments have shown that the ejecta are concentrated in a thin cone with central axis at 10° from the target surface. The fragments are mainly solid. The average size is larger than the fragments produced in classical oblique impacts. The number of hypervelocity fragments is dramatically increased.

4 Implementation in SPENVIS

For the particle/wall interaction model two versions are implemented in SPENVIS: one based on the ESABASE user manual which corresponds to the above descriptions and one on the Handbook for Designing MMOD Protection The following meteoroid and debris models are implemented in SPENVIS:

References

Anderson, B. J., Review of Meteoroids/Orbital Debris Environment, NASA SSP 30425, Revision A, 1991.

Cour-Palais, B. G., Meteoroid Environment Model-1969, NASA SP-8013, 1969.

Drolshagen G., and J. Borde, ESABASE/Debris Meteoroid/Debris Impact Analysis, Technical Description, Issue 1 for ESABASE version 90.1, 1992.

Grün, E., H. A. Zook, H. Fechtig, and R. H. Giese, Collisional Balance of the Meteoritic Complex, Icarus, 62, 244-272, 1985.

Hechler, M., Collisional probabilities at geosynchronous altitudes, Adv. Space Res., 5, 47-57, 1985.

Jenniskens, P., Meteor Stream Activity I, The annual streams, J. Astron. Astrophys., 287, 990-1013, 1994.

Kessler, D. J., Space Station program definition and requirements, JSC 30000

Kessler, D. J., J. Zhang, M. J. Matney, P. Eichler, R. C. Reynolds, P. D. Anz-Meador, and E. G. Stansbery, A Computer-Based Orbital Debris Environment Model for Spacecraft Design and Observations in Low-Earth Orbit, NASA Technical Memorandum 104825, 1996.

Lemcke, C., Enhanced debris/micrometeoroid environment models and 3D software tools, Final Report of ESTEC Contract No. 11540/95/NL/JG, 1998.

Lemcke, C., G. Scheifele, and M.-C. Maag, Enhanced debris/micrometeoroid environment models and 3D software tools, Technical Description of ESTEC Contract No. 11540/95/NL/JG, 1998.

MBB-ERNO, Standard Methods for Analysis of Micrometeoroid and Orbital Debris Impacts (MDPS Design Guideline), COL-MBER-000-RP-0022-03, 1988.

McBride, N., and J. A. M. McDonnell, Characterisation of Sporadic Meteoroids for Modelling, UNISPACE KENT, 1996.

McCormick, B., Collision probabilities in geosynchronous orbit and techniques to control the environment, Adv. Space Res., 6, 119-126, 1986.

Oswald, M. et al., Upgrade of the MASTER model,Final Report to ESOC Contract 18014/03/D/HK(SC), 2006.

Sdunnus, H., Meteoroid and Space Debris Terrestrial Environment Reference Model "MASTER", Final Report to ESOC Contract 10453/93/D/CS, 1995.

Taylor, A. D., The Harvard Radio Meteor Project meteor velocity distribution reappraised, Icarus, 116, 154-158, 1995.

Taylor, A. D., W. J. Baggaley, and D. I. Steel, Discovery of interstellar dust entering the Earth's atmosphere, Nature, 380, 323-325, 1996.


This page is based on Drolshagen and Borde [1992], Lemcke [1998], and Lemcke et al. [1998].

Last update: Mon, 12 Mar 2018