Chapter 8

Orbital Operations Safety

Tommaso Sgobba and Firooz A. Allahdadi

Chapter Outline

8.1 Space Situational Awareness Systems and Space Traffic Control

Fernand Alby

Space Situational Awareness Systems

Introduction

The objective of space surveillance is to create an inventory of objects above a certain size that are in orbit around Earth. This inventory (catalog) provides information about the origin of the objects (name, launch country) and their trajectory (orbital parameters), allowing them to be located at a later date. To fulfil this objective, different types of sensors must be used: first of all, detection systems with a wide field of view to see passing objects above a certain size and roughly calculate their orbit in order to locate them later on; then, tracking systems with a narrow field of view are used to follow a specific object, in order to take trajectographic measurements and improve knowledge of its trajectory. Basically, these detection and tracking systems are radars for low-orbiting objects or telescopes for objects in higher orbits. They can be located either on the ground or in orbit. The population of smaller debris below the size limit is discerned in a statistical and no longer deterministic way: this information is given by flux models such as Ordem (NASA) or Master (ESA). Some authors consider that space surveillance also includes functions relating to space weather and the monitoring of near-Earth asteroids whose trajectories could threaten Earth. In the rest of this article, we will only be considering the functions related to space traffic control, that is to say detecting and tracking artificial objects in orbit around the Earth.

The main source of information is the space surveillance system set up by the United States, which provides the most comprehensive information. Russia has a similar system for which very little information is available. On the other hand, the International Scientific Optical Network (ISON) telescope network provides a detailed catalog of objects in geostationary orbit. Finally, France has limited capability with the GRAVES system.

USA

(See NASA Handbook 8719.14, 2008; Chatters & Crothers, 2009.)

The US space surveillance network (SSN) is a combination of optical and radar sensors used to support the Joint Space Operations Center’s (JSpOC) mission to detect, track, identify, and catalog all manmade objects orbiting the Earth. The JSpOC, located at the Vandenberg Air Forces Base, is in charge of programming the sensors, and recovering and analysing the data to compile and manage the catalog. The American network carries out the following functions:

• detects new man-made objects in space;

• produces a running catalog of man-made space objects;

• determines which country is responsible for an orbiting or re-entering space object;

• charts the present position of space objects and plots their anticipated trajectories;

• predicts when and where a space object will re-enter the Earth’s atmosphere.

Several types of sensors are used to carry out these functions:

• Phased-array radars (PAR) can maintain tracks on multiple satellites simultaneously and scan large areas of space in a fraction of a second. These radars have no moving mechanical parts to limit the speed of the radar scan – the radar energy is steered electronically. In a PAR there are many thousands of small transmit/receive antennas placed on the side or face of a large wedge-shaped structure.

• Conventional radars use moveable tracking antennas or fixed detection and tracking antennas. A detection antenna transmits radar energy into space in the shape of a large fan. When a satellite intersects the fan, energy is reflected back to the detection antenna, where the location of the satellite is computed. A tracking antenna steers a narrow beam of energy toward a satellite and uses the returned energy to compute the location of the satellite and to follow the satellite’s motion to collect more data.

• Electro-optical sensors consist of telescopes linked to video cameras and computers. The video cameras feed their space pictures into a nearby computer that drives a display scope. The image is transposed into electrical impulses and stored on magnetic media. Thus, the image can be recorded and analyzed in real-time or later.

• The Midcourse Space Experiment (MSX) satellite is an Earth-orbiting satellite with a payload containing a variety of sensors, from ultra-violet to very-long-wave infrared. Originally a platform for Ballistic Missile Defense Organization projects, the MSX was moved to the SSN in 1998. Since that time, the visible-light Space-Based Visible (SBV) sensor on MSX has contributed significantly to the tracking of geosynchronous objects and is serving as a pathfinder for a future space-based space surveillance system (SBSS).

The SSN sensors are divided into three categories: dedicated, collateral, and contributing. A dedicated sensor is a US Strategic Command (USSTRATCOM) operationally controlled sensor with a primary mission of space surveillance support. A collateral sensor is a USSTRATCOM operationally controlled sensor with a primary mission other than space surveillance (usually, the site’s secondary mission is to provide surveillance support). Contributing sensors are those owned and operated by other agencies that provide space surveillance support upon request from the JSpOC.

The SSN is able to follow objects of around 10 cm in low orbit (possibly 5 cm for low-altitude objects and those on inclined orbits) and objects of around 1 m in geostationary orbit. On 1 April 2011, the public catalog contained 14,870 objects.

Combined, these types of sensors make up to 100,000 satellite observations each day. This enormous amount of data comes from SSN sites such as Maui, Hawaii; Eglin, Florida; Thule, Greenland; and Diego Garcia, Indian Ocean. The data is transmitted directly to JSpOC via satellite, ground wire, microwave and phone. Every available means of communication is used to ensure a backup is readily available if necessary (Figure 8.1.1).

image

FIGURE 8.1.1 Configuration of the US SSN in 2006.

To set up and manage the catalog, the SSN uses two different orbitography models: a model called General Perturbations (GP), an analytical model, which uses a simplified representation of the forces and a Special Perturbations (SP) model based on numerical integration with a more precise representation of the forces. Only less precise information produced with the GP model is available on the Space Track website. This information is given in Two-Line Element (TLE) format: the SSN and Committee on Space Research (COSPAR) number of the object, average orbital parameters on two lines (see detail and format on the Space Track website, https://www.space-track.org/).

The TLE only provides a rough knowledge of the orbit: the uncertainties can reach several kilometers from the creation of the TLE and then quickly deteriorate over time. In particular, this information is not precise enough to carry out a reliable collision risk prediction. Nevertheless, since 2010, the JSpOC has been setting up agreements with operators wishing to be provided with collision warnings based on precise data (SP) and messages called Conjunction Summary Messages (CSM) are sent to the operators. CSMs contain precise information on dangerously close objects and the associated uncertainties, which makes it possible to calculate the collision probability.

Russia

Space surveillance system

Very little information is available on the Russian Space Surveillance System. According to a Jane’s report (Jane’s Space Systems and Industry, April 27, 2007), Russia’s SSS Space Surveillance System relies primarily on Dnepr and Darayl-UM radars operating in the very high frequency (VHF) range near 150 MHz at eight sites. The Russian sites are: Irkutsk, Murmansk and Pechora in Russia; Sevastopol and Uzhgorod in Ukraine; Balkhash in Kazakhstan; Mingechaur in Azerbaijan; and Riga in Latvia (Riga’s Darayl-UM was scheduled for 1995 demolition). Dnepr should continue through 1998. Another site at Baranovichi was under construction, but its status is unclear. Operation of many of the space surveillance and telemetry, tracking and command (TT&C) sites is now erratic, with facilities dropping out and then returning to service. Several key installations have been shut down and remaining assets consolidated around primary sites in Russia mentioned above. C-band and higher frequency radars are beginning to be acknowledged. Observations are also provided to the SSS by two ABMD Anti-Ballistic Missile defence radars in the Moscow region near Sofrino and Chekhov operating at ultra-high frequency (UHF) near 400 MHz. This combined network generates some 50,000 observations daily to maintain a catalog of nearly 5000 objects, most in low Earth orbit (LEO). Owing to geographical limitations, Earth satellites at low inclinations are difficult or impossible to track. As a whole, the SSS radar sensors appear to be limited to a range of about 4000 km.

International Scientific Optical Network (ISON)

Since it started in 2005, the ISON project (ISON Worldwide Scientific Optical Network et al., April 2009) has grown considerably: today it brings together 25 observatories located in nine countries (Bolivia, Georgia, Moldova, Russia, Spain-ESA, Switzerland, Tajikistan, Ukraine and Uzbekistan). It makes use of 32 telescopes with aperture diameters between 0.2 and 2.6 m. The Keldish Institute of Applied Mathematics (KIAM) of the Russian Academy of Sciences provides overall coordination of the network.

The ISON network is the only civilian non-governmental network capable of providing space surveillance information on high altitude orbits. The quality of the information provided is at least equal to that of the United States Space Surveillance Network.

The system covers the whole of the area around the geostationary orbit and is able to detect and track objects in this area as well as on eccentric orbits of high altitude (HEO – high elliptical orbits, GTO – geostationary transfer orbits or Molnyia-type orbits). See Figure 8.1.2.

image

FIGURE 8.1.2 The telescopes of the ISON network.

Hundreds of hitherto unknown objects have been discovered in geostationary Earth orbit (GTO) and HEO thanks to ISON.

The number of telescopes should gradually be increased to 40: this network will allow every object over 1 m and around 90% of objects over 50 cm located in the vicinity of the geostationary orbit to be constantly and independently monitored.

France

Since 2005, France has had a limited low-orbit space surveillance capability using the GRAVES system (Grand Réseau Adapté à la Veille Spatiale). The system consists of a single bistatic radar associated with significant calculation systems and a secure transmission network.

The transmitter site is located near Dijon: it consists of four transmission antennas, each covering a sector of 45° in azimuth and 20° in elevation: the total cover in azimuth is 180°. The radar continuously transmits in VHF.

The receiving site, located near Apt, consists of 100 antennas distributed over a metallic disc. Each antenna is linked up to an individual receiver and the detection beam is then digitized (Figure 8.1.3).

image

FIGURE 8.1.3 Emitting and receiving antennas of the GRAVES system.

The system is operated by the French Air Force. It enables a catalog of around 2500 objects located below 1000 km in altitude to be independently compiled and maintained. Every object is seen at least once every 24 hours and the orbital parameters are determined using a single passage.

Transmitting antennas emit a continuous low-frequency signal towards a given angular section of space. The receiving site, located nearly 400 km away, houses a large number of omnidirectional antennas. Based on the elementary signals picked up by these antennas, a narrow-lobe beam is produced. The direction of this lobe provides an angular measurement of the object detected, while the frequency shift between the emitted signals and the received signals measures its radial velocity.

Based on this brand-new concept, the GRAVES radar provides angular and radial velocity measurements. These are fed into the orbital processing algorithms developed by ONERA (Office National d’Études et de Recherches Aérospatiales) to calculate the orbital parameters of the detected satellites (Figure 8.1.4).

image

FIGURE 8.1.4 Principle of the GRAVES radar.

ESA

In 2008, the ESA Member States decided to set up a space surveillance program (SSA – Space Situational Awareness), which aims to equip Europe with independent systems 10 years from now (Declaration of Preparatory Space Surveillance Programme, 2010). In the field of space surveillance, services cover the detection and tracking of objects over a given size and the establishment of a catalog of these objects, the launch of collision risk warnings, the recommendation of avoidance maneuvers, as well as the detection of explosions in orbit. Services will also cover the prediction of high-risk re-entries. In the field of space weather, services involve monitoring the Sun, solar wind, radiation belts, the magnetosphere and the ionosphere. Furthermore, preliminary activities concerning near-Earth objects and establishing a catalog of these objects will be carried out.

The preparatory phase of this programme covers the 2009–2012 timeframe and includes four elements:

• The core elements.

• Space weather activities.

• The pre-development of critical radar sub-systems and the assembling of functional models (radar elements).

• The pilot data centers.

The operational development phase will be decided at the end of the preparatory phase at an Agency Council meeting at ministerial level, decided an additional preparatory phase in order to consolidate the definition.

Space Traffic Control

The International Academy of Astronautics (IAA) Cosmic Study on Space Traffic Management defines traffic management as: “The set of technical and regulatory provisions for promoting safe access into outer space, operations in outer space and return from outer space to Earth free from physical and radio-frequency interference” (Cosmic Study on Space Traffic Management, 2006).

The main objective of space traffic control is therefore to avoid the risk of collision in orbit by maneuvering satellites. However, it should not be forgotten that objects in orbit are subject to the laws of astrodynamics: they move very quickly (8 km/s) in low orbit and their trajectory can be only very slightly altered by maneuvering, for those whose propulsion system is in good working order. For example, altering the inclination of a satellite located 800 km away by 1 degree requires a delta V of about 140 m/s which corresponds to a 100 kg propellant consumption in the case of a classical 2T satellite when considering a 280 s Specific Impulse. Consequently, the most frequent maneuvers take place in the orbital “plane”: an alteration of the semi-major axis and/or eccentricity, which leads to an alteration in the orbital period and therefore makes the object pass a given point a little earlier or later to avoid a collision. In-plane maneuvers may also be implemented to ensure a separation in the radial direction between both objects.

Space traffic control cannot therefore be compared to traffic control in other fields, in particular air traffic control. For example, it is not possible to allocate navigation corridors to some objects, as is done in air space. For instance, an object in transfer orbit sweeps across all the altitudes between its perigee in low orbit and its apogee near the geostationary orbit. Furthermore, orbital planes turn under the effect of the Earth oblateness, and at different speeds depending on their semi-major axis and their inclination. In space, objects circulate therefore at all altitudes and in all directions.

In practice, collision risk management can be carried out in a preventive way by adopting coordination procedures between satellite operators, or in a systematic way by monitoring the collision risks of every satellite with all other cataloged objects (prediction).

Prevention of Collision Risk

Following the increase in the number of objects in space and the heightened risk for satellites, preventive measures (mitigation) have been decided, first of all at the level of the main space agencies, then at international level within the framework of the Inter-Agency Space Debris Coordination Committee (IADC), which brings together 12 agencies: ASI (Italian Space Agency), CNES (French Space Agency), CNSA (China National Space Administration), CSA (Canadian Space Agency), DLR (German Space Agency), ESA (European Space Agency), ISRO (Indian Space Research Organisation), JAXA (Japan Aerospace Exploration Agency), NASA (US National Aeronautics and Space Administration), NSAU (State Space Agency of Ukraine), ROSCOSMOS (Russian Federal Space Agency), and UKSA (UK Space Agency). So, in 2002, the IADC published the document “IADC Space Debris Mitigation Guidelines”, which is still the reference in this field today. This subject was also discussed within the more general framework of the United Nations, at the Committee on Peaceful Uses of Outer Space (COPUOS). This committee was created in 1959 and brings together 70 countries: its work resulted in the publication in 2007 of the document UN-COPUOS Mitigation Guidelines, which defines the main principles of the preventive measures to apply in space regarding space debris. These documents recognize, in particular, the need to urgently protect two areas in space because of their congestion and their importance for space applications, namely the low Earth orbit (LEO), which corresponds to altitudes below 2000 km, and the GEO region, which corresponds to altitudes between ± 200 km around the geostationary orbit and to inclinations between ± 15 degrees (Figure 8.1.5).

image

FIGURE 8.1.5 Protected regions.

At the same time, the International Organization for Standardization (ISO) is developing standards for space debris from these reference documents, which manufacturers and operators will be able to use.

These texts are not compulsory in nature. They are only recommendations. The applicable rules depend on each country and are presented in laws (e.g., Space Operations Act in France) or by licensing systems as in the United States or the United Kingdom. These different texts remain consistent for they are all derived from the IADC document.

Low earth orbit
Management of a constellation

The Iridium constellation (Figure 8.1.6) consists of 66 operational satellites plus a few satellites on hold as replacements in case of breakdown. The satellites are distributed over six orbital planes with an inclined orbit of 86.4° and at an altitude of 780 km (six satellites in every plane).

image

FIGURE 8.1.6 The Iridium constellation.

The orbital planes intersect in the vicinity of the poles: all the constellation satellites go through these two points. There is therefore a potential risk of collision at these points. Moreover, a collision between two constellation satellites would produce two debris clouds situated in neighboring orbits, which in turn would come and cross the orbits of the other satellites, representing an increased risk of collision. To avoid collision risks between satellites in the same constellation, every satellite’s position in its orbit must be constantly controlled in order to ensure a safe distance with satellites on neighboring planes when they pass close to the poles.

Protection of operational orbit during end-of-life maneuvers

When several satellites are located in the same orbit (in the case of the SPOT observation satellites: four satellites in the same plane inclined at 98.5° and at an altitude of 825 km), the operations to withdraw one of the satellites from service lead to its drifting in relation to the other satellites of the same family and must be carried out while taking account of the collision risks. See Figure 8.1.7.

image

FIGURE 8.1.7 SPOT satellites operational orbit.

In the case of SPOT1, the objective of the end-of-life operations in November 2003 was to reduce the altitude of the perigee so as to ensure the satellite re-entry into the atmosphere in less than 25 years. To do so, nine successive maneuvers were performed at the apogee, which allowed the perigee altitude to be lowered to 580 km. However, prior to that, it was necessary to control the exit from operational orbit: two additional maneuvers were added at the beginning of the sequence to carry out a classic Hohman transfer from the circular orbit of 825 km in altitude towards a circular orbit in the same plane 15 km lower. Any collision risk thus having been removed, it was then possible to set the series of deorbiting maneuvers in motion without disturbing the other satellites. The sequence of operations was the following:

Phase 1: Exit from operational orbit and transfer to 15 km lower.

Phase 2: Eight apogee maneuvers enabling the perigee to be gradually lowered.

Phase 3: Final maneuver to completely empty the tanks and lower the perigee.

Geostationary orbit

Traffic control is particularly important in GEO because of the high number of satellites present, not only because of their movement but also because of the consequences of a possible collision: the debris created would spread throughout the orbit and would permanently remain there in the absence of natural disturbances such as atmospheric drag in low orbit.

In geostationary orbit, the space surveillance data available is imprecise and not frequently updated. As it is not easy to predict collision risks in this area, operators must avoid crossing the operational area and should coordinate with operators in charge of neighboring positions.

The protected geostationary area is the area of altitude between GEO −200 km and GEO + 200 km (see Figure 8.1.5). It includes the station-keeping window area (± 37.5 km around the geostationary altitude; see Figure 8.1.8) and a navigation corridor below and above the window area, used when the satellites are moving around (Figure 8.1.9).

image

FIGURE 8.1.8 Typical station-keeping window.

image

FIGURE 8.1.9 Station-keeping zone and navigation corridor.

In geostationary orbit, traffic control is performed on several levels:

• Orbital position control.

• During the final positioning operations when the satellite drifts to meet up with its station window.

• During the positioning phase to control the presence of several satellites in the same window.

• During position changes when the satellite goes from one station longitude to another.

• At the start of end-of-life operations when the satellite leaves its station-keeping post to join the graveyard orbit.

Management of longitudes and frequencies

The geostationary orbit is a unique resource used by many satellites: its parameters must satisfy very precise conditions (circular orbit in the equatorial plane and at an altitude of 35,786 km) to have a fixed position in relation to the Earth. On the other hand, more and more operators want to place satellites on this orbit mainly for telecommunications applications. Strict control of this orbit is therefore necessary to avoid interference between satellites and collision risks. These two aspects are controlled by the International Telecommunication Union (ITU), which is a specialized United Nations agency based in Geneva. The ITU includes 192 State Members as well as more than 700 members of the private sector and academic institutions.

An orbital position and a frequency slot are allocated to every operator. Orbital positions are given in the form of a station longitude, for example Telecom 2D to 8°W. The satellite must remain near this longitude, in practice in a station-keeping window whose typical values are ± 0.1 degrees longitude and ± 0.1 degrees latitude. The non-recovery of station-keeping windows ensures the absence of collision risk between neighboring satellites, except at the time of arrival in the window (positioning) or exit (service withdrawal operations): see Injection, longitude changing and disposal operations below.

Coordinated station-keeping of several satellites in the same window

Certain longitudes are particularly sought after (coverage of Europe, the Americas, or Asia for example). Some operators can then choose to place several satellites in the same station-keeping window (in the case of Astra; see Figure 8.1.10) or else several operators can choose to share the same window: coordinated station-keeping should then be carried out for these satellites in order to control their relative movements under the effect of natural disturbances and station-keeping maneuvers to avoid any collision risk.

image

FIGURE 8.1.10 The seven Astra satellites observed with the Tarot telescope.

By using the orbital parameters of each satellite, it is possible to ensure that their natural movement in the window is such that a safe distance will be retained between them (separation in longitude or in eccentricity and/or inclination. Combined separation strategies in eccentricity and inclination are also feasible, as shown in Colocation Strategy and Collision, November 1989, for the satellites TDF1, TDF2, TVSAT2 and Olympus at 19°W.

Injection, longitude changing and disposal operations

In theory, operational geostationary satellites remain in their station-keeping window to avoid collision risks. Problems can arise in the case of loss of control of a satellite that is going to start drifting and cross its neighbors’ windows. Coordination between operators is then necessary to control this risk, at least until the orbital parameters of the drifting object have not been estimated and made available to the other operators via available catalogs.

The problem also arises when a satellite arrives in its window (positioning), or has to modify its longitude, or departs during service withdrawal operations: in these cases, the orbit-changing maneuvers should take into account the resulting drift between the maneuvers and the risks of windows being crossed in which other controlled satellites are located.

• Injection: at the time of the positioning phase it is preferable to aim for a lower apogee altitude underneath the operational station-keeping area. For the final transfer towards the station-keeping window the same precautions as for the position changes must be taken into account: limiting the eccentricity drift in order not to interfere with the operational area, coordinating with the immediate neighbors for the final phase.

• Longitude changing: when a satellite has to change station position (longitude), the semi-major axis of its orbit must be altered, which causes a drift towards the East or West depending on the change indication of the semi-major axis. To avoid crossing the other station-keeping windows and to reduce the collision risk, the drift must be carried out in the navigation corridor described above. The eccentricity of the drift orbit leads to a daily variation in altitude (see Figures 8.1.11 and 8.1.12). Therefore, the eccentricity has to remain weak so as not to interfere with the operational area.

image

FIGURE 8.1.11 Drift orbit with high eccentricity.

image

FIGURE 8.1.12 Drift orbit with low eccentricity.

• Disposal operations: for end-of-life operations the problem is similar to the one faced in positioning: the window exit is made “upwards” to then reach the graveyard orbit located 200 km above the geostationary orbit and beyond. Maneuvers must take into account satellite presence in neighboring orbital positions: the initial drift phase must be carried out with a weak eccentricity. Information from the other operators is necessary, as well as coordination to deal with unforeseen events such as a thrust transfer.

In any case, coordination with operators that have satellites in the vicinity must be set up from the beginning of the mission: in fact, during positioning, station-keeping, longitude changing, or end-of-life maneuvers, one of the scheduled sequence maneuvers may not be carried out. The resulting drift can lead the satellite to return to a neighboring satellite’s window. Before carrying out the operations, contact points must be defined, as well as procedures and data exchange formats. During the operations it is important to exchange precise orbital parameters with the neighbors as well as the scheduled maneuvers. Finally, it is recommended that the distance with other satellites be constantly monitored and that a warning be generated when this distance falls below a predefined threshold (Recommended Practices for Traffic Management, 2008).

Prediction of Collision Risk

In-orbit collision risk

The number of objects in space has steadily increased over the past several years, so predicting collision risk in orbit has become one of the main tasks for control centers in charge of monitoring and handling satellites. The risk of losing a satellite during a collision is no longer negligible, as is shown in Table 8.1.1 for two satellites in low orbit (Space Debris Models and Risk Analysis, 2004).

Table 8.1.1

Annual collision risk

Objects > 10 cm Cataloged objects
ENVISAT 0.015 0.0073
ERS2 0.0039 0.0021

It should be noted that a collision would not only result in the destruction of the two objects, but also the creation of a large amount of debris. For example, in the case of the collision between the Iridium 33 and Cosmos 2251 satellites, two clouds of debris were created: 349 debris cataloged from the Iridium 33 satellite and 809 debris cataloged from Cosmos 2251 (situation in June 2009).

Operators manage this risk using the available space surveillance data. This data enables them to predict when objects will get dangerously close to each other several days in advance, calculate the risk, and carry out a collision avoidance maneuver, which slightly alters the satellite’s trajectory so it passes a safe distance away from the dangerous object.

The monitoring procedure is relatively onerous because of the imprecision of the available data: it generally consists of a basic level of automatic monitoring, which detects potential risks, and which then have to be analysed in more detail by experts in orbitography. If the risk appears to be serious, trajectography measurements are requested from the available radar systems (generally military systems): a better knowledge of the trajectory of the dangerous object can be obtained from these measurements and then the decision to carry out a collision avoidance maneuver can be taken if necessary. The whole prediction procedure takes several days (usually three). Finally, it should be noted that the collision avoidance maneuver alters the trajectory of the monitored satellite and generally makes it necessary to interrupt the satellite mission, which can be a serious constraint in the case of an observation satellite. A return maneuver into the nominal orbit will then be necessary before resuming the mission. All of this, therefore, mobilises important resources: experts, controllers, radars, calculation systems, telemetry/telecommand (TM/TC) stations, etc. and also results in propellant being consumed, which accordingly reduces the lifespan of the satellite. To reduce the impact of these maneuvers, it is sometimes possible to anticipate a scheduled maneuver (a station-keeping maneuver, for example) by performing earlier than planned a maneuver that it would have been necessary to do in any case, thereby reducing the cost of propellant.

By way of example, here is the collision risk monitoring report carried out at CNES in 2010. Eighteen satellites were involved in this monitoring: 353 risks were identified by the automatic process with a collision probability of over 10−4. At the same time, 92 warnings were received from the JSpOC. Analysis of these cases resulted in 21 requests for radar measurements or support being made to the JSpOC (collision probability of over 10−3) and in the end, 13 collision avoidance maneuvers were performed.

Risk collision at launch

During the launch phase and the initial orbits, the upper stage of the launcher and the satellites put into orbit will cross the orbits used by other operators: this is particularly true in the case of a geostationary transfer orbit whose perigee is in low orbit and whose apogee borders on 36,000 km in altitude. These newly injected objects will not be listed in the catalogs until some hours later (generally 48 hours), which means that other space users have no means of monitoring the collision risks between these new objects and their satellites. This is particularly important in the case of vehicles with a crew on board (International Space Station (ISS) for example) whose control center cannot monitor the risk posed by these objects.

The launch operator alone has access to the information on the planned trajectory: he/she can therefore carry out the collision risk prediction. This prediction must be made for any launch date within the launch window while taking account of all the objects put into orbit (launcher stage, satellites, structural elements) over a period of around 48 hours (see Figure 8.1.13). If there is a risk, shifting the launch time by a few seconds ensures a safe distance between objects. Once 48 hours have passed it can be assumed that the new objects have been cataloged and that each operator is able to perform his own monitoring.

image

FIGURE 8.1.13 Risk collision between the ISS and a launch trajectory from Kourou: each launch time corresponds to a different trajectory because of the Earth’s rotation.

The main difficulty in predicting collision risk comes from including the dispersions of the different objects on the orbital parameters at injection: the propagation of these dispersions over 48 hours leads to relatively large volumes of error around each object and could lead to total closure of the launch window if all the cataloged objects were considered in the analysis. It is for this reason that collision risk prediction at launch is generally limited to manned vehicles or some satellites of special interest.

Conclusion

The main objective of space traffic control is collision avoidance. To accomplish this, three conditions must be met:

• Space users must have access to reliable and precise information on the trajectories of all objects above a certain size: today this function is fulfilled by the JSpOC, whose policy of data dissemination was developed after the collision between Iridium 33 and Cosmos 2251. The other sources of information are not as comprehensive and their availability is not guaranteed.

• A code of conduct must exist: the work carried out within the framework of COPUOS and IADC resulted in recommendations that are applied voluntarily by most operators. These recommendations consist of preventive measures, particularly relating to end-of-life management, and operational measures aiming to predict objects coming dangerously close to each other in orbit or at launch. The application of these measures is now reinforced by their integration into national legal systems such as licensing systems or laws on space operations.

• Increased coordination is necessary between operators sharing neighboring orbital positions. This is particularly the case in geostationary orbits where a station longitude is allocated to every satellite by the ITU. Every operator therefore knows its “neighbors” and can inform them of maneuvers for positioning, longitude change or end-of-life exit from the area. This information is also particularly important during operational life, as a breakdown may lead to a satellite drifting towards positions occupied by others.

Today this system works relatively well: for example, it can be noted that between 2000 and 2010 the number of GEO operators correctly performing end-of-life operations rose from 27% to 69%. Of course, there are still operators that do not apply any of these measures, and at present there is no coercive mechanism to force them to comply with these rules or any financial penalty as is the case in other fields on Earth or at sea.

References

1. Chatters EP, Crothers BJ. Space surveillance network, Chapter 19, Space Primer: Air Command and Staff College Space Research Seminars. 2009; Available at http://space.au.af.mil/au-18-2009/; 2009; Accessed February 2013.

2. Colocation Strategy and Collision Avoidance for the Geostationary Satellites at 19 degrees West, M.C. Eckstein, C.K. Rajasingh, P. Blumer, presented at the CNES international symposium on Space Dynamics, Toulouse (France), 6–10 November 1989.

3. Cosmic Study on Space Traffic Management, report published in 2006 by the International Academy of Astronautics (IAA).

4. Declaration of Preparatory Space Surveillance Programme, ESA/Program Board-SSA (2010)29.

5. ISON Worldwide Scientific Optical Network, I. Molotov et al., 5th European Conference on Space Debris, Darmstadt (Germany), 30 March–2 April 2009.

6. Jane’s Space Systems and Industry, April 27, 2007, section Space Defence-Operators.

7. NASA Handbook 8719.14 Handbook for limiting orbital debris. 2008; Available at http://www.hq.nasa.gov/office/codeq/doctree/hb871914.htm; 2008; Accessed February 2013.

8. Recommended Practices for Traffic Management in the GEO Protected Region, L.Lorda, Geo end-of-life workshop CNES 2008.

9. Klinkrad H. Space debris models and risk analysis. Berlin: Springer-Verlag; 2004.

8.2 Orbit Design for Safety

David Finkleman

The objective of this subchapter is to suggest criteria for designing satellite and constellation orbits with on-orbit safety as a consideration. This is the most essential element of space traffic safety management.

Fortunately, there have been few attributable collisions among satellites (Iridium Incident Highlights, 2009). Unfortunately, there are thousands of relatively close approaches daily among objects in Earth orbit. The consequences of collisions and their likelihood have justified significant investment in avoiding them, and in diminishing the population of potential collision threats. This can be accomplished with orbit architectures that avoid highly populated regimes and collision risks within a mission system, constellation, or formation while still meeting mission needs.

This subchapter reviews the collision risks of current orbits and constellations, describes constellations and orbit architectures conceived for minimum collision risk, and suggests schemes for prevention of potential collision in orbit and constellation design.

This subchapter emphasizes aspects of orbit and constellation design principles well covered in texts (Wertz & Larson, 1999) and regulations (NASA Procedural Requirements, 2009; NASA STD, 2009). Wertz and Larson (1999) is without much argument unique in citing rules for constellation design, describing the engineering trade-offs, and including specifically collision avoidance and debris mitigation as design considerations.

In the current and evolving near-Earth space environment, safety should not be an afterthought. It can be among the principal constraints without compromising mission capability. There may be additional cost.

Distribution of Objects in Earth Orbit

The non-uniform distribution of the Earth’s mass and irregularities in the influences of the Sun, the Moon, and other massive celestial bodies are significant benefits as well major complications. Sun-synchronous orbits would not exist if the Earth’s gravitation were uniform. Halo orbits about libration points exist only in three-body or more complex gravitational fields. Rapid, energy efficient transfer to geostationary orbits must consider the influence of the Moon. Innovative deep space missions that acquire energy from gravity assisted planetary flybys would be incomprehensible in a simple, Newtonian two-body (satellite–Sun) approximation.

Perturbations complicate mission planning. They cause orbits to rotate around the Earth’s axis (westward nodal regression), and the axes of the orbital ellipse to rotate around the normal to the orbit plane (precession of the line of apsides), (Fortescue et al., 2003). Virtually all orbits, even geostationary orbits, require station-keeping maneuvers to maintain satellites in the desired relationship to the Earth or other significant celestial references. To a practical approximation, these phenomena are mitigated in specific orbits. For example, nodes do not regress appreciably in strictly polar orbits, and the lines of apsides do not precess if a specific relationship exists between orbit altitude and inclination (generally retrograde).

These phenomena cause satellites for important missions to be concentrated in a few orbit regimes, notably Sun-synchronous, geostationary, and orbits critically inclined to mitigate apsidal precession. The semi-synchronous regime (halfway to geostationary) is also very desirable for precise positioning, timing, and navigation satellites, and it offers a good compromise between transmission power requirements for ubiquitous presence and coverage footprint on the Earth. Orbit inclinations corresponding to the latitudes of major launch sites are also relatively crowded, since launching eastwards accrues the greatest delta V increment from Earth rotation.

Figure 8.2.1 is the distribution of satellites in Earth orbit displayed as an apogee–perigee graph known as a Gabbard Plot. Circular orbits have unit slope. Apogee is by definition greater than perigee; therefore there are no orbits above the 45 degree line. The dense orbital regimes are obvious. Note that this depiction neglects orbit inclination. Figure 8.2.2 is a comparable statement of orbit distribution according to inclination and apogee. Critical inclinations are apparent.

image

FIGURE 8.2.1 Apogee-perigee distribution of resident space objects.

image

FIGURE 8.2.2 Inclination distribution of resident space objects. (Oltrogge and Kelso, 2011)

The frequency of close approaches among satellites is another measure of orbit crowding. Although the minimum separation between pairs of satellites is a convenient discriminator, it is not a sufficient measure of risk since orbits are known precisely, the probability of physical contact is low (Finkleman, 2009). With that qualification, it is worrisome that there are daily thousands of encounters with closest approach less than 5 km and relative velocities of many kilometers/second. On 30 August 2011, 17,695 approaches closer than 5 km among unclassified, cataloged objects were estimated over the ensuing seven days (http://celestrak.com/SOCRATES).

Using familiar and conveniently disposed orbits naturally leads to crowding. Perhaps “convenient” is too great an abstraction. The orbits most used, such as geostationary, are much more than just convenient. However, the figures above demonstrate that there are orbital regimes that are still relatively free of collisions. We will demonstrate that many of these are only marginally less “convenient,” if at all.

Rules of the Road

There are fundamental physical reasons why traffic control in space cannot be analogous to traffic control for automobiles, ships, or aircraft. The momenta and inertia of satellites are many orders of magnitude larger than those for the other vehicles. Gravitation may be a very small force in orbit. The actual force required to change a satellite’s trajectory sufficient for collision avoidance is also relatively small, particularly in geostationary orbit. However, the energy, i.e. fuel required to responsively maneuver, is large compared to the station-keeping requirements. A single maneuver on short notice might reduce on orbit lifetime considerably.

The best collision avoidance approach is to coordinate satellite operations in advance, but this might be inefficient or even infeasible. Satellite locations determine coverage and the continuity of customer service. Commercial operators are reluctant to share maneuver plans with each other. It might be unwise to coordinate maneuvers for national security reasons. There might also be incontrovertible mission requirements that could place other satellites in jeopardy.

Courteous rules of the road as exist on highways and sea lanes are inapplicable in space. The motivation for having a satellite transcends comity. Satellite operators are disaffected, enjoying freedom of space that is independent of any sovereign authority or sanctions. Often one of the conjunction partners is uncontrollable debris.

Passive Safe Operation Alternatives

If courteous behavior is unlikely and responsive avoidance is infeasible, the next alternative is to operate satellites in orbits that are unlikely to suffer crowding and collision risk. This might compromise essential mission capability. Either more satellites might be required for the same capability or additional propellant, propulsion, and guidance systems will be required.

This is the approach used in astronomy. Given comparable atmospheric clarity (seeing), telescopes must be located far from terrestrial sources of illumination (light pollution). Radio telescopes must be able to discriminate among extremely weak receptions from distant galaxies. They mitigate electromagnetic interference by being located remote from human electromagnetic activity. This makes these operations less efficient and more expensive.

We will explore several alternatives for safe operation, minimizing fratricidal risks and vulnerabilities to debris. For given mission requirements, we will characterize compromises to mission execution and the additional costs of more assured safety. We intend this to be the basis for incorporating safety as an architectural consideration.

Inherent collision risk is the most important safety discriminant. Our ability to know where satellites are continuously and precisely is another consideration. Orbits and the satellites themselves should also consider how well and how often they can be monitored. Space surveillance in any embodiment will not be able to see everything, everywhere, all of the time. More and better distributed sensors will only push the threshold of perception farther, not eliminate it. As sensors proliferate the value added by each new one will diminish.

We can mitigate the diminishing return of more sensors by designing satellites, orbits, and constellations with observability as a consideration. Useful highly eccentric, highly inclined orbits are least observable where their orbits are likely to change most, during rapid perigee passage in the far Southern Hemisphere.

Ideally, increasing observability could have no impact on mission capability or satellite control. Dilectis and Mortari pioneer analytic techniques in orbit design for observation from designated Earth sites (Dilectis & Mortari, 2011). Satellite signatures can also be augmented for observability. Many close approaches could be eliminated with better and more responsive trajectory data that would make possible very small mitigation maneuvers.

Observability is much more than line of sight to a satellite from a terrestrial sensor or even the duration of access. The number of independent observation opportunities, the geometry of the access, the frequency of revisits, and measurement quality matter greatly. Vallado and Griesbach have examined the effect on orbit determination precision of these satellite–sensor pair characteristics (Vallado & Griesbach, 2011). They have exquisitely cataloged hundreds of possible trajectory observation sites. Harrison and Finkleman used the same techniques to estimate the potential space situational awareness contributions of the numerous sensor systems (predominantly telescopes) that might participate word-wide (Finkleman, 2010).

Many dimensions of capability must be considered, for example whether a sensor is active (radar or laser) or passive (telescope), whether the satellite must be illuminated by the Sun or Moon, what on-site communications and data processing there are, and whether satellite signatures are augmented actively or passively. Vallado and Griesbach codify the constraints. The range of orbits that is feasible with adequate observability is potentially large enough that a greater variety of orbits could be employed for the sake of mitigating crowding and collision probability.

Figure 8.2.3 shows the extent of inclination of low Earth orbit satellites that are accessible with reasonable frequency and duration using modest telescopes with active laser illumination. These all contribute to precise orbits and ephemerides (POE) that others use to calibrate their instruments and orbit estimation techniques.

image

FIGURE 8.2.3 Representative distribution of sensors defining sufficiently observable orbit regimes. (From Vallado & Griesbach)

Characteristics of Unsafe Orbits

There are some obvious safety guidelines. First, an inclined orbit will cross the planes of all satellites with lower inclination twice on each revolution. Even satellites at the same altitude in different planes at different inclinations whose orbital locations (true anomalies) are out of phase will eventually come close together because of orbit perturbations. The higher the inclination, the more orbit planes of lower inclination satellites the subject satellite will cross. If the satellite belongs to a constellation all members of which are at the same altitude and inclination, the separation between orbits decreases with increasing latitude. The orbits will be clustered densely near Poles. The more satellites in each plane, the higher the frequency of close approaches.

Despite the unarguably greater encounter risk, highly inclined orbits are very desireable. Highly inclined orbits are a necessity for operators that must serve users at high latitudes. Critical inclination with minimal precession of line of apsides is a relatively high inclination.

Iridium is an example of a high risk constellation. On 30 August 2011, more than 1000 approaches within 5 km were estimated with rocket bodies, other debris, and even other Iridium satellites over the ensuing week. Much of the debris from the Cosmos 2251–Iridium 33 collision remains in Iridium-like orbits, increasing the encounter risk to other Iridium satellites. Iridium satellites themselves and the cross-linked communication network are extremely robust to on-orbit losses, but that might be a necessity.

Alternative Orbits for Surveillance

John Draim was principal inventor of the elliptical Ellipso, Virgo, and Cobra constellations and holds many patents in space technology. He also holds the only patents that provide nearly continuous global coverage with only four satellites (Draim, 1987). Circular orbit arrays for continuous global coverage require a minimum of five satellites. Using elliptical constellations, coverage can be concentrated geographically by latitude and/or longitude, and/or by time of day. Circular constellations generally waste much coverage on areas of little interest. Draim has received international recognition for his work on elliptical constellations that can provide users with more efficient system designs using the minimum number of satellites. His most recent “Droplet” designs allow the use of fixed ground antennas (similar to GEO systems) pointed toward mid- and high-latitudes and thus avoid any radio frequency (RF) interference with the GEO systems. In addition, they are designed to avoid major orbital debris fields, as well as the infamous van Allen radiation belt.

There are no general analytical techniques for constellation design. Selecting satellite orbits for a specific mission is a trade-off among: coverage, station-keeping effort, launch effort, onboard power requirements, communication access and energy requirements, persistence, and several other factors. There are no unique or universally optimal solutions. Numerous combinations of orbit characteristics, numbers of satellites, and onboard communication or surveillance capabilities will satisfy almost any specific mission criterion. We add avoiding close approaches with other satellites and minimizing the consequences of fragmentation to the trade-off criteria.

The first example is Draim’s eight satellite droplet constellation (Draim, 2009). Through numerical optimization, Draim arrived at constellation families for complete coverage of the Northern Hemisphere. The families include constellations of eight, 12, and 16 satellites, with 63.435 degree inclination, 16 hour periods, and two day repeating ground tracks. Figure 8.2.4 is the eight satellite configuration.

image

FIGURE 8.2.4 Draim eight satellite, 16 hour period, constellation for complete and continuous Northern Hemisphere coverage.

This constellation provides near stationary presence at high latitudes with very small antenna pointing excursions.

Figure 8.2.5 demonstrates that the satellite laser ranging network alone has complete visibility of the constellation at apogee. Darker areas are access by a single sensor, lightest areas are triple or more coverage. Figure 8.2.6 demonstrates complete coverage at perigee.

image

FIGURE 8.2.5 Satellite laser ranging network coverage at constellation apogee. Darkest regions have at least single coverage. Lightest regions have greater than triple coverage.

image

FIGURE 8.2.6 SLR coverage at constellation perigee altitude.

Figure 8.2.6 shows that the droplet orbits are visible to one or more single-lens reflex (SLR) sensors at perigee altitude. There is a small region off the tip of South America which enjoys only single or double coverage.

Conjunction searches against the complete, openly available satellite catalog confirm almost complete freedom from close approaches. Over weeks, there are only two approaches within 100 km of any of the Droplet Constellation satellites.

This is a representative example of an exquisitely safe constellation that fulfills a well defined mission. In some ways, this constellation has unique advantages. For example, such continuous coverage of a mid-latitude region anywhere in longitude would require many more satellites in more conventional orbits, particularly circular.

There are several additional practical and cost trade-off elements. In general, more satellites might be necessary than in more conventional architectures. These include capabilities of each satellite in terms of aperture required for desired resolution, motion required for reasonable relative collaborations such as frequency difference of arrival discrimination, and launch costs.

Conclusion

We have reviewed some considerations for safe orbit and constellation architectures. We have shown that the feasible volume of orbital space that minimizes collision risk is many times greater than currently “convenient” orbits. We have also shown that safe operation will entail additional cost, for example, requiring more satellites. However some, such as Draim’s patented tetrahedron, actually require fewer satellites for the whole Earth coverage mission and diminish collision risk. Most of the orbits noted in this paper actually require less orbital energy than many “convenient” orbits. However, achieving those orbits might be more energy intensive and difficult. Those orbits might also require more station-keeping energy and suffer greater perturbation-induced perturbations. Much research is required, and engineering practices must be extended to achieve the goal.

References

1. Dilectis F, Mortari D. Analytic Orbit Design for Earth Sites Observation. AIAA Paper 11-147 Girdwood, AK: AAS/AIAA Astrodynamics Specialist Conference; 2011.

2. Draim JE. A Common-Period Four-Satellite Continuous Global Coverage Constellation. Journal of Guidance. 1984;Vol. 10.

3. Draim JE. Sixteen-hour droplet constellations for Northern Hemisphere coverage. Heriot-Watt, Edinburgh, Scotland: The 27th AIAA International Communications Satellite Systems Conference, ICSSC 2009, Edinburgh Conference Centre; 1-4 June 2009.

4. Finkleman D. Discriminating Threatening Conjunctions with Data Fusion Principles, 2009. Pittsburgh, Pennsylvania: AAS/AIAA Astrodynamics Specialist Conference; 2009; AAS 09-409.

5. Finkleman D. Space and Verification. Technical Assessment, Eisenhower Center for Space and Defense Studies. Vol. II 2010.

6. Fortescue P, Stark J, Swinerd J. Spacecraft Systems Engineering. (3rd ed.) Wiley & Sons. 2003.

7. Iridium Incident Highlights. Growing Risk of On-Orbit Collisions, Beck Iannotta. Space News 19 Feb 2009.

8. NASA Procedural Requirements for Limiting Orbital Debris, NASA Office of Safety and Mission Assurance. (May 2009). NPR 8715.6A, Change 1.

9. NASA STD 8719.14 Process for Limiting Orbital Debris. (September 2009), Change 4.

10. Oltrogge DL, Kelso TS. Getting to Know Our Space Population from the Public Catalog, AAS 11-416, 2011. Girdwood, AK: AAS/AIAA Astrodynamics Specialist Conference; Aug 2011.

11. Wertz JR, Larson WJ. Space mission analysis and design. 3rd ed. Dordrecht, The Netherlands: Kluwer; 1999.

12. Vallado DA, Griesbach JD. Simulating Space Surveillance Networks, AIAA Paper 11-580. Girdwood, AK: AAS/AIAA Astrodynamics Specialist Conference; Aug 2011.

8.3 Conjunction Analysis

Erwin Mooij and Ron Noomen

Introduction

In June 2011, the six-person crew of the International Space Station (ISS) reportedly took refuge in the two Soyuz capsules that were docked to the station (NASA, 2011). They did so to prepare for a possible impact by a tracked but uncataloged debris object, which could potentially hit the ISS with a relative velocity of 39,000 km/h. In the end it passed at a distance of 725 m. It does not take much imagination to understand what the consequences of such a collision would be. Many questions immediately pop up, such as: Where does this threat come from? Can we predict an impact, and if so: how? What are the chances of a real hit? How often does this occur?

The problem is directly related to the presence of man-made objects in space: space debris. An illustration of the situation is given in Figure 8.3.1, which gives a snapshot of the debris objects in the direct vicinity of the Earth. All objects move with a velocity of about 28,000 km/h, and the ISS has to be navigated through this mess in a safe way. The current situation is very different from the situation when Sputnik was launched (October 1957), or when Yuri Gagarin made the first-ever manned flight (April 1961): in 2011 the observable space debris totaled some 16,000 objects, and the number is growing.

image

FIGURE 8.3.1 Illustration of orbit debris in low Earth orbit. Image source: NASA.

This subchapter will deal with this problem of possible collisions in space. More precisely, it will focus on the techniques that can be applied to assess the collision risk for an arbitrary active satellite (from now on called “user satellite”). Other elements to deal with this space debris problem, such as collision avoidance techniques, protection and removal, will not be discussed here. Also, this chapter will address the risk induced by man-made space debris only; (micro)meteoroids, i.e. natural elements in the space environment, pose a similar threat, but this is at a much lower level than that induced by man-made objects. The reader is referred to textbooks such as that by Klinkrad, (2006) for more information on these topics.

In the next section, a brief overview of the space debris population will be given, and the section after that illustrates the problem in more detail. The next discusses analytical techniques to make a first-order assessment of possible collisions, and there follows a description of techniques to deal with the uncertainties in the orbits of the objects involved in a numerical fashion. Finally, the last section gives the conclusions.

Debris Population

Our knowledge of the population of space debris is based on measurements. Traditionally, these are taken by observing stations on Earth, which employ two different techniques: radar and optical. The former is based on reflections of pulses of energy, initially emitted by ground stations. Since the efficiency of this technique drops off with a factor 1/r4 (where “r” is the distance between the ground station and the object in space), its application is limited to objects of about 2 mm diameter in low Earth orbit (LEO) for the most capable stations. However, regular observations are required to correlate these to an individual object and determine its orbit, and this limits the minimum diameter of catalog objects to about 10 cm (again, for LEO). At high altitudes, such as the geostationary orbit (GEO), information on the whereabouts of objects is collected by observing sunlight reflections; here the minimum debris diameter of catalog objects is about 30 cm. Stations routinely tracking space debris are operated by organizations in the USA, Russia, Europe and Japan.

As an alternative, and in order to better quantify and understand the (development of the) space debris population, measurements can also be done in space, i.e. from a platform that itself orbits Earth. Classical contributions from this perspective have come from the post-mission analysis of impacts on NASA’s Long Duration Exposure Facility (LDEF; in space from 1984 until 1990), of impact craters in the windows of the Space Shuttle Orbiters (which flew 135 missions in the period 1981–2011, typically with a duration of 2 weeks each), and of other retrieved spacecraft and/or parts thereof. Examples of more novel in-orbit instruments that provide observations on a continuous basis are Russia’s Geostationary Orbit Impact Detector (GORID; operational at GEO altitude from 1996 until 2002) and ESA’s Debris In-Orbit Evaluator (DEBIE; attached to the PROBA-1 satellite in LEO and operational from 2001 until now, and to the ISS, 2008–now). Both GORID and DEBIE use particle impact detection techniques to observe and quantify incoming debris objects (which must be limited to small scales for obvious reasons). Clear advantages of such modern in-situ observations are their continuity and the reduction of the minimum size for detection: GORID can measure objects with a mass down to 1014 g (Klinkrad, 2006). Recently (September 2010), the US launched the first “Pathfinder” satellite of the Space-Based Space Surveillance System (SBSS) to a Sun-synchronous orbit at 630 km altitude.

Figure 8.3.2 gives an overview of the amount of space debris throughout the years. This is one of the most complete overviews, based on observations taken by the US Space Surveillance Network (SSN) and converted into a catalog by USSTRATCOM (2012). The plot also includes active satellites. For obvious reasons, the picture does not show the full number of objects orbiting Earth: first, because of the limitations on particle size as mentioned above, and second, it is restricted to public information (and so military objects and the remnants thereof are not included). It does show that the number of objects is steadily increasing: In 2011, it encompassed a total of almost 16,000 objects, as diverse as intact satellites, spent rockets stages, and fragments thereof. Sharp increases, with about 2000 observable objects each, are related to the Chinese anti-satellite test (January 2007) and the collision between the operational Iridium-33 and the obsolete military Cosmos-2251 (February 2009).

image

FIGURE 8.3.2 The number and origin of catalogued space debris objects, as a function of time. Image source: NASA.

Extrapolating today’s conditions (number, strategies) to the year 2100, continuing business-as-usual (i.e., similar launch rates, no cleanup programs, increased chances of collisions and explosions) will lead to a total of about 50,000 LEO objects of 10 cm or more in size (Klinkrad, 2006). This is an increase with respect to the current situation of 160%, and it will be clear that this provides bad prospects for the future of spaceflight.

The real situation in space is worse than depicted in Figure 8.3.1. Figure 8.3.2 shows that there is a multitude of objects with size below the catalog threshold (i.e., smaller than 10 cm and 30 cm in LEO and GEO, respectively). It is estimated that about 600,000 objects larger than 1 cm and some 300 million objects larger than 1 mm are orbiting Earth at present (Klinkrad, 2009), outnumbering the detectable objects by far. If these populations follow the same trend as predicted for the objects of 10 cm or more, the space debris situation would develop into something really dramatic. Some analysts even claim that spaceflight itself is at stake: If no precautions are made to prevent the further build-up of debris, the risk for space missions by the end of the century will have become so high that it will be basically not acceptable to fly any more missions.

Figure 8.3.1 showed an illustration of the space debris population for the LEO environment (the picture could easily be expanded to also include the GEO environment). This picture suggests that the situation is completely random. This is indeed the case for the positions of the debris objects at an arbitrary epoch, but their behavior with time does exhibit structure. Similar to operational spacecraft, the space debris objects obey the laws of nature, and as a consequence orbit Earth to first order in so-called Kepler orbits. Such orbits can be fully characterized by a set of six constants, the Kepler elements. This gives the opportunity to predict the course of individual objects and to quantify the risk of collisions, which will be the topic of the next sections.

Problem Definition

The previous section briefly discussed a number of aspects of the space debris population, including the trend of increasing numbers. Based on these numbers, one can make a first-order estimate of the chances of being hit by another object, when designing a new mission (Wertz et al., 2011):

image (1)

In this equation, PC represents the probability of collisions (–), SPD is the spatial density of the debris population (#/m3), AC is the cross-sectional area of the user satellite (m2), T is the mission lifetime (s), and VREL is the relative velocity of the impacting particle with respect to the user satellite (m/s). As an example: at an altitude of 800 km, the average concentration of observable debris objects is about 6 × 1017/m3. At this altitude, the circular velocity is 7450 m/s; a representative relative velocity is a factor √2 larger for the situation when hitting at right angles. So, for a new mission with a cross-sectional area of 10 m2 and a mission lifetime of 5 years, the chances of being hit by an object of 10 cm or more would be about 0.1%. In the same manner, one can also (statistically) estimate the chances of being hit by a particle larger than 1 cm; here, the concentration at 800 km is about 2 × 1015/m3, and the chances of being hit would become 3.3%. Similarly, the chance of a hit by particles larger than 1 mm would become 100%.

In a similar fashion, or by looking at how the trajectories of individual objects evolve, one can make predictions of the total number of collisions between any pair of objects, for the coming century. It is claimed that by the year 2100, the cumulative number of collisions will have grown to about 50 (Klinkrad, 2006).

It is obvious that we do not want a repeat of the Iridium–Cosmos collision (actually, this was the first fatal incident involving an operational satellite; the non-fatal collision between the radio-amateur satellite Cérise and an explosion fragment of a spent Ariane upper stage, in July 1996, was the first collision between any two tracked objects, with a severed gravity boom of the former as a result). Therefore, the population of objects in space (operational and debris) has to be monitored continuously, and actions have to be taken when an impact is imminent. Typically, such an action is to change the orbit of one of the satellites involved such that the risk of a collision is effectively reduced to zero. This is done by a so-called evasive maneuver, where a rocket engine burns a certain amount of propellant to achieve a large enough velocity/orbit change.

Obviously, this can only be done when at least one of the vehicles involved is still under control, and it comes at the expense of propellant mass usage; the latter of course has adverse effects on the operational lifetime of the vehicle. So, such maneuvers are usually refrained from unless there is a compelling reason not to lose the vehicle, for example, it concerns a manned mission (ISS, manned transfer vehicle) or an extremely valuable asset (ENVISAT or the Hubble Space Telescope, having cost about 2.5 billion euros and 6 billion dollars, respectively). Triggered by warnings generated by various surveillance organizations, the operators of the ISS perform evasive maneuvers about once per year, on average (NASA, 2011). For other missions, operators typically “take their chances,” if only to keep the nominal lifetime of the mission as long as possible. This is claimed to be the case with the collision between the Iridium and Cosmos vehicles (Wertz et al., 2011). The techniques to predict such problems are dealt with in the subsequent sections.

First-Order Risk Assessment

To assess the chances of a possible collision between the user satellite and any other object in space, one should study the trajectories of all spacecraft and check whether the user satellite and any other of these objects would come too close. Such an activity is typically done for a week ahead, and is of course restricted to the objects for which orbital information is available, i.e. objects that are present in the catalog(s) mentioned in the previous sections. When the objects approach one another within a certain distance, for example 10 km, one speaks of a so-called conjunction, which might trigger an evasive maneuver or another action. The process described here is called a conjunction analysis, and is done continuously to monitor the environmental dangers to manned and high-value spacecraft.

Although the fundamentals of such an analysis are very straightforward, it is a process that comprises a number of complications. First of all, the satellite positions and velocities at the initial (reference) epoch are not perfectly known, but have an inherent uncertainty. This is most generally described by a 6-dimensional covariance matrix, including (squares of) formal uncertainties of individual components (Cartesian position and velocity) and correlations between these parameters. When making a prediction of the future state of any vehicle, one should in principle also predict the behavior of this covariance matrix, which takes quite some cpu time. In addition, the current database contains about 16,000 objects. When interested in potential conjunctions between all pairs of objects, evaluating all N × (N–1)/2 options with a fine enough time step becomes an extremely cumbersome process (here, “N” represents the number of objects in the catalog, and since LEO satellites move with an absolute velocity of more than 7 km/s, possible conjunction events have to be predicted with a resolution of better than 1 s). Even when the predictions are restricted to pairs involving the user satellite only, it still takes huge amounts of cpu time. Third, this prediction process has to be reliable enough to not only miss possible conjunctions, but also not to generate so-called false alarms.

In an effort to overcome such practical problems, this prediction analysis is typically performed in a number of successive steps. First, the conjunction analysis is restricted to a propagation of the best-known estimate of the state-vector of the satellites at epoch, without regarding their inherent uncertainties, and the limits for a possible conjunction are set relatively high (e.g., 25 km). In the next phase, a more detailed analysis is done, now also involving a propagation of the uncertainty of the positions and velocities at epoch. Since this next step is much more demanding in terms of cpu-time, this process is typically performed for a limited set of spacecraft pairs only. Both steps will be discussed below in more detail: the first one in this section and the more detailed one under “Detailed Risk Assessment” below.

To relieve spacecraft operators, the Center for Space Standards and Innovation (CSSI) runs a service called SOCRATES (Satellite Orbital Conjunction Reports Assessing Threatening Encounters in Space) that identifies potential conjunctions on a regular basis (CSSI, 2011). Operators are free to use this information according to their best insight.

TLEs and State-Vector Propagation

The propagation of the state-vector of any satellite starts with the selection of the most appropriate set of information at a reference epoch. This information can consist of the 3-dimensional position and velocity in Cartesian components, or be given as Kepler elements; the exact form is unimportant for the purpose of this conjunction analysis. For obvious reasons, one should use a state-vector that is most reliable, and that is defined and available in a way that is common for all spacecraft being analyzed. The well-known sets of Two Line Elements (TLEs) generated by USSTRATCOM (2012) satisfy both requirements: they are typically given with frequencies ranging between several times per day for manned spacecraft, to several times per week for “average” active satellites, to once per (few) week(s) for extinct rocket boosters and such. The state-vector information in each TLE is given on two successive lines of ASCII code, which provide information on the Kepler elements, derivatives thereof, and a few other parameters related to the dynamic behavior of the vehicle under consideration, such as an atmospheric drag scaling parameter. The TLEs are made available by the US authorities (USSTRATCOM, 2012) and CelesTrak (Kelso, 2010). As mentioned before, the database encompasses about 16,000 objects at the time of writing, and it is the most complete and consistent set of data publicly available.

In principle, the propagation of a satellite state-vector can be done in a number of ways. The most common way is to numerically integrate the position and velocity of a satellite, while evaluating the instantaneous accelerations that act on the vehicle for each individual time-step. Such an integration can be done with Runge–Kutta (RK) integration schemes, where the step-size (fixed or variable) and the accuracy of the specific technique (RK4(5), RK7(8)) can be chosen at will (typically, this is a trade-off between accuracy and efficiency). Alternatives for RK methods are also possible, of course. Another option is to integrate the Kepler elements that describe the position and velocity of the vehicle in a geometric way. In that case, one uses the so-called Lagrange Planetary Equations, or the Gauss form thereof (e.g., Cornelisse et al., 1979). Here again, the integration type and specifics can be chosen at will.

However, all of this loses its relevance when one chooses to use the TLEs. Since these elements essentially provide the Kepler elements, the values of which were estimated in an orbit determination with an environment model with limited accuracy (central gravity, a few low-order terms of the gravity field, possible resonance, aerodynamic drag, solar radiation pressure, third-body acceleration due to Sun and Moon), one should stick to the conventions of this force model when making extrapolations of these state-vector solutions. Fortunately, such a propagation can easily be done, since USSTRATCOM and CelesTrak provide propagators ready for this task. In the past, and depending on the problem at hand, the user had to choose between the SGP4, SDP4, SGP8 and SDP8 propagators (SGP and SDP are abbreviations for simplified general perturbations and simplified deep-space perturbations, respectively) (Hoots & Roehrich, 1980). However, through time improvements have been made to these propagators, and currently the best version publicly available is the SGP4 propagator as described by Vallado et al. (2006); it inherently covers al possible orbital regimes (i.e., low and high orbits).

Brute Force

In the most simple conjunction analysis, one follows a so-called brute-force approach: propagate the pair of any two vehicles to a certain epoch, evaluate whether they approach one another too close or not, and then move on to the next epoch (e.g., 1 s further in time). Once the entire time interval of interest has been covered, the user switches to the next pair of vehicles. Ideally, the test “coming too close” suffices to deal with the volume physically taken by the (combination of) spacecraft, i.e. distances of several tens of meters. In reality, one has to incorporate uncertainties in the representation of the satellite orbit, which has repercussions for the techniques described here and the subsequent section.

It will be clear that this approach is not very efficient, if only for the fact that in this way the state-vector of a particular (user) satellite is integrated repeatedly (which can easily be solved, of course). Also, the relative velocity between an arbitrary pair of vehicles can be several tens of km/s, so when tested against a minimum distance of, e.g., 25 km at specific epochs, care must be taken not to miss any potential conjunction (or worse). To prevent this from happening, one typically increases the critical distance Rcr to a threshold value Rth, which allows for the relative motion in between evaluated steps Δt (Figure 8.3.3).

image

FIGURE 8.3.3 Definition of the critical radius Rcr and threshold radius Rth, defined by the maximum relative velocity 2Vesc of the two objects.

Knowing that the individual objects move with an absolute velocity, which cannot be larger than the local escape velocity, one can readily derive the following equation for this threshold radius):

image (2)

One further refinement is to increase this distance with the effect of possible accelerations during each evaluation step. As a worst case, one can assume that the acceleration of an individual vehicle is the gravity acceleration at sea level, so the relative acceleration is twice this value. The new expression then becomes:

image (3)

The expressions above show one of the main problems with this approach: in order to identify a reasonable amount of possible conjunctions, the testing criterion Rth cannot be set too high, i.e. the evaluation time-intervals of this brute-force method cannot be set too large. As an example, if Rcr and Δt are set at 25 km and 1 s, respectively, one would have to test against an Rth of 36 km. This implies the computation and pair-wise comparison of 86,400 times 16,000 objects for a single day (which will be very demanding on the computers involved). Instead, switching to an evaluation step-size of 10 s (60 s) would result in a value for Rth of 138 km (732 km). Although the number of evaluations decreases by factors 10 and 60, respectively, the larger value for Rth will significantly increase the number of possible conjunctions (which all need further in-depth analysis).

Clearly, this process can be optimized. Without going into detail, an optimal step-size appears to be about 20 s; optimal in the sense of reasonable number of conjunctions for a reasonable cpu-time. Yet, an analysis of a full catalog for a single day easily costs about a full day of cpu-time. To make this first-order assessment of possible conjunctions more efficient, filters and sieves are typically applied.

Filters and Sieves: Theory

Following the brute-force method, one would have to analyze all pairs of state-vectors and determine their mutual distance for all epochs in the dataset. As indicated, this would be an extremely laborious process. So, it would be very advantageous to have a tool to eliminate certain combinations of satellites directly, without going into the computation of distance, velocity, and such, for each individual evaluation epoch. Tools have been developed for this purpose: filters and sieves. In essence, they do the same thing: a quick yet reliable elimination of “impossible combinations”; filters do this for a complete interval of interest (e.g., a day), whereas sieves do this for an individual epoch.

Filters

The best-known filters are perigee–apogee, geometrical, time and Taylor series.

Figure 8.3.4 gives an illustration of the perigee–apogee filter. It is based on a simple evaluation of the perigee and apogee distances of any pair of vehicles. Essentially, it filters out pairs for which the highest perigee q exceeds the lowest apogee Q by a distance D:

image (4)

image

FIGURE 8.3.4 Illustration of different cases for the perigee–apogee filter.

Referring to Figure 8.3.4, the combination (1,2) can be ruled out for further analysis (when the margin D has been chosen appropriately: for the entire day), whereas the combinations (1,3) and (2,3) may lead to conjunction “hits”. This perigee–apogee filter is an extremely efficient filter, typically rejecting about 60% of possible pairs beforehand.

A second filter considers the 3-dimensional relative geometry of the two Kepler orbits (Figure 8.3.5).

image

FIGURE 8.3.5 The closest approach distances d1 and d2 between two elliptical orbits.

If the two closest distances d1 and d2 both exceed the criterion D, this pair is also filtered out. The exact details of this evaluation can be found in the paper by Hoots et al. (1984); this reference also deals with special geometries, like (near)coplanar orbits.

Obviously, the characteristics of satellite orbits (size, orientation) are uncoupled from the actual position of a vehicle in such an orbit. Therefore, one may also eliminate possible pairs for conjunction analysis when their positions in the orbits are too far off (think of a pair of geostationary satellites, positioned on mutual sides of the Earth: orbital size and orientation are perfectly identical, but they will never approach one another). The crossing points which have been identified with the geometrical filter just described, must be connected to the times of passage by each of the two vehicles, and then the observation can be made whether a real conjunction is imminent or not. Again, with a fairly limited number of computational steps one can possibly filter out a pair of objects for a relative long period of time. It will be clear though that both the geometrical and the time filter are much more demanding than the straightforward perigee–apogee filter!

Finally, one can also predict the epoch of closest approach and the corresponding distance, using a Taylor-series approach:

image (5)

Here, image, image and image are the relative (3D) position, velocity and acceleration at an (arbitrary) evaluation epoch t0; the 3D position x is evaluated at an arbitrary other epoch t. Truncating the series expansion to the linear term, the first-order estimate of the epoch of conjunction tconj becomes:

image (6)

and the distance dconj:

image (7)

If needed, one can do a series of iterations such that the final result is accurate to a fraction of a second; in practice, and of course assuming a reasonable step-size for the evaluations, one iteration will be sufficient. With this specific filter, a large number of individual evaluation points can be skipped for in-depth analysis.

Sieves

Alternatively, a number of tools have been developed that do a similar thing: eliminate a certain combination of spacecraft without going into (relatively) lengthy computations. Now, the conclusion only holds for the particular epoch under consideration. These techniques are called sieves rather than filters. Almost all sieves are based on the comparison of coordinates.

X

The simplest implementation of this is to compare just one coordinate, such as the difference in x-coordinates of the pair under consideration at the particular epoch (“rx” in the equation below). If it exceeds a certain boundary value of, for example, 25 km, one can disregard this particular occasion readily, without the need to do any other calculation. Experience tells that this is a good approach, and when applied after a perigee–apogee filter has been applied already, it will still filter out 90–99% of all occurrences (depending on the evaluation step-size). Moreover, by virtue of the absence of any computations (the satellite positions have been computed earlier), it is a very fast technique.

image (8)

Y

In a similar fashion, one can test the difference in y-coordinates (“ry”). It comes as no surprise that the performance of this filter is similar to that of the sieve for x-coordinates, both in rejection percentage and in cpu-time.

Z

The obvious third possibility is to evaluate the difference in z-coordinates. Again, this sieve has the same performance as the previous two.

r2

Rather than looking at individual components of coordinates differences, it is also possible to compare the radii of the two spacecraft positions for the particular epoch: when this difference exceeds a certain threshold, it will not pose the risk of a collision. Obviously, rather than comparing the radii, it makes more sense to compare the squares of these values, without any loss of quality. After all, these distances are in principle computed by adding the squares of the differences in each Cartesian component, and not going through the effort to take the square root of each value makes the sieve more efficient. This filter also eliminates 90–99% of all possible pairs, and needs slightly more cpu-time than the X/Y/Z sieves.

image (9)

Minimum distance

It is possible to refine the test on distance by deriving the minimum distance and testing this against a stricter test criterion Racc. The idea behind this sieve is depicted in Figure 8.3.6, which shows the minimum distance between the two objects extrapolated from the position and velocity values at epoch t0.

image

FIGURE 8.3.6 The minimum distance between two objects rmin is dependent on the distance r0 and relative velocity V0 at the evaluation epoch t0.

The minimum distance is computed with the following expression:

image (10)

Since the relative velocity of the two objects is already incorporated in the value of rmin, one can use a more simple test criterion Racc that only represents the critical volume and the relative (worst case) acceleration:

image (11)

Clearly, this sieve involves more computations (so is slower in terms of cpu-time), but is stricter, so in essence will reject more pairs.

Fine r2

In this sieve, the actual relative velocity is used rather than the maximum (escape) velocity of a satellite. Again, the main gain comes from working with a stricter test criterion (at the cost of more computations):

image (12)

with

image (13)

Fine conjunction

The fine conjunction sieve is based on the numerical evaluation of the null values for Vapp2 to accurately find the point of closets approach. Clearly, this test is the strictest one, but involves the largest number of evaluations (i.e., cpu-time).

Combination

In practice, one tries to identify the potential conjunctions with a minimum of computational effort (before going into a detailed analysis with uncertainties, see next section). Clearly, the sieves as presented above increase in distinguishing performance, but at the cost of more cpu-time. Therefore, it comes as no surprise that the most efficient conjunction analysis is based on a succession of filters and sieves: perigee–apogee, r2, minimum distance, fine r2, fine conjunction. As a matter of fact, the X/Y/Z sieves (although very simple) do not provide much benefit over the more complete, 3-dimensional r2 test, and are therefore eliminated.

Detailed Risk Assessment

In the last section, first-order assessment gave us those combinations of satellites and/or debris that can potentially collide. For this subset a detailed analysis should be carried out. Simple orbital-state models such as those based on Kepler elements cannot be used for perturbed-orbit calculations with inherent uncertainties. Complex models are required that can in principle only be solved numerically (see “Perturbed-Orbit Propagation” below). Extensions to the aforementioned simple models will, in general, not solve the singularities for, for instance, circular orbits, and also lead to a complex system of equations. Possible alternatives to these models are those based on Cartesian coordinates, modified equinoctial elements and the unified state model. The latter two have excellent numerical properties in terms of computational speed and accuracy.

To assess the risk involved with satellite collisions accurate measurements of the initial position and velocity are important to guarantee a reliable prediction. Orbits are generally unstable in the along-track direction and the larger the measurement error the larger the error will be after a certain period of time. To account for this behavior one usually does many simulations in the form of a Monte-Carlo analysis, although the outcome can only give a certain probability whether a collision will occur (see “Orbital Stability and Monte-Carlo Analysis” below). A novel, alternative, method is that of verified interval propagation, which propagates the uncertainty related to the initial conditions and can give with a mathematical certainty of 100% that no collision will occur. Alternatively, it can lead to the conclusion that a collision cannot be ruled out. This method will be illustrated on basis of the Cosmos–Iridium collision in 2009 (see “Verified Interval Propagation” below).

Perturbed-Orbit Propagation

A simplified representation of the orbital state does not suffice for accurate orbit prediction. The unperturbed Kepler model may serve well as a first-order approximation or as a means to analytically analyze the orbit-propagation problem. It will not hold, though, when predictions of the satellite motion under the influence of not only the gravitational attraction of the main body but also a variety of perturbing accelerations are required. These perturbations may come from the Earth’s atmosphere or magnetic field, the third-body attraction of the Moon or the Sun, or the solar-radiation pressure, to name but a few.

In general-purpose astrodynamics tools, the state of an orbiting body and the related equations of motion, are often described in Cartesian coordinates. Its form is very simple, and if we separate the external acceleration into a main and perturbing component, the equations read:

image (14)

where rI is the position vector in the inertial frame, μ the standard gravitational parameter of the central body, and apert,I the sum of perturbing accelerations.

For numerical integration, eq. (14) can be written as a system of first-order differential equations:

image (15)

with

image (16)

These Cartesian coordinates have a physical meaning (position and velocity) and the orthogonality of the system results in simple equations of motion and transformations. It is therefore the preferred state model in many celestial-mechanics applications. However, an unperturbed two-body problem is an orbit, the shape and orientation of which can be described by five constants, and the position along the orbit by a single variable. The Cartesian model does not use this information, requiring a smaller step-size to reduce the integration errors.

The two-body orbit-shape and orientation description is the basis for the traditional Kepler elements. These elements define the shape and orientation of the satellite orbit and its position along this orbit (usually given by the true anomaly, θ). Such a state model, including the differential (or variational) equations to account for the change in orbit parameters due to the perturbations, is often more efficient and/or accurate in numerical integration. This is a result of the larger step-size that can be used without the loss of accuracy, because the main gravitational force has no or a linear effect on the variation of the orbital elements. This is in contrast with the Cartesian velocity components that are all driven by the main force with a quadratic dependency on the position. The major disadvantages of using the perturbed Kepler elements are the extra complexity of the model that make it harder to implement and verify, and singularities for both circular, equatorial and polar orbits.

To account for these singularities the modified equinoctial element (MEE) set, a mathematical combination of the Kepler elements, has been defined (Walker et al., 1985):

image (17)

image (18)

image (19)

image (20)

image (21)

image (22)

where the traditional Kepler elements are given by: a, the semi-major axis; e, the eccentricity; ω, the argument of periapsis; Ω, the right ascension of the ascending node; i, the inclination; and θ, the true anomaly. All but L are constant for an unperturbed two-body problem. When perturbation forces act on the body, the following differential equations describe the change of the MEE due to these forces (Betts & Erb, 2003):

image (23)

image (24)

image (25)

image (26)

image (27)

image (28)

In the above equations ae1, ae2 and ae1 are the three components of the perturbing acceleration vector, expressed in the radial direction, e1, perpendicular to the radius vector, e2, and a component perpendicular to the orbital plane, e3. As models for the perturbing accelerations are commonly available in an inertial, Cartesian frame, ae can easily be obtained from apert,I through a frame transformation involving ω, Ω, i, and θ.

The MEE has three disadvantages over Kepler elements. First, it is obvious that the MEE lose the physical meaning that Kepler elements have. Second, the variational equations are more complex, resulting in a higher computational cost. Third, the rotation matrix in MEE is more complex than the Euler angles in the Kepler elements (this transformation matrix is required to transform the perturbing forces from the inertial to the orbital frame).

Despite these disadvantages, the major advantage of being singularity free is of prime importance. Of course, many more orbital-state models and related variational equations have been developed over the years. The MEE has been compared with many of them and came out as the best one in terms of accuracy and computational speed (Hintz, 2008).

A second orbital-state model that shows a potential of being a fast and accurate orbit propagator is the unified state model (Vittaldev et al., 2010, 2012). The unified state model (USM) uses seven state variables to describe the state of an orbiting body: three to define the shape of the orbit in the velocity phase space and four quaternions to define the orientation of the orbital plane with respect to the inertial reference frame.

C, Rf1 and Rf2 are functions of the radial and angular momentum that determine the energy of the satellite and the size and position of the orbit in velocity phase space (the velocity hodograph); see Figure 8.3.7. Rf1 and Rf2 are the components of R in a convenient intermediate reference frame that is rotated about the line of nodes. R has a constant direction, and points in the direction of the velocity at periapsis. C is always perpendicular to the radius vector. The satellite’s velocity, v, is the sum of both, i.e., v = R + C.

image

FIGURE 8.3.7 USM velocity components C and R together with the radial and angular velocity components ve1 and ve2 shown in the velocity (left) and position (right) phase space of a Kepler orbit.

For an unperturbed two-body problem, the magnitude of both R and C is constant. Since the orbit in velocity phase space is constant for an unperturbed two-body problem, these variables are also constant for the unperturbed case. Where the MEE use the shape and size of an orbit (ellipse) in the position phase space, the USM uses C and R to define the orbit (circle) in the velocity phase space.

The USM uses a quaternion to keep track of the orientation of the orbital plane (Altman, 1972). q4, q1, q2, q3 are the four quaternion elements, image, and are also called Euler parameters. Since four parameters are used to define three degrees of freedom, the following constraint on the quaternions needs to be introduced: image.

Quaternions have two major advantages over Euler angles. No singular conditions occur in a rotation, the degrees of freedom are never lost, i.e., there is no gimbal-lock, and all angular functions are algebraic instead of trigonometric (Kuipers, 1999).

As mentioned, the unified state variables C, Rf1 and Rf2 are constant for the unperturbed two-body problem. If perturbing forces act on the body, the equations of motion in terms of the perturbing force components in the e1, e2 and e3 direction read:

image (29)

image (30)

where:

image (31)

which becomes singular for q3 = q4 = 0. This corresponds with retrograde equatorial orbits (i = 180°). This means that the USM breaks down for retrograde equatorial orbits. This orbit can be avoided by reformulating it with a negative orbital velocity and zero inclination.

The body-fixed velocity components ve1 and ve2 are given by:

image (32)

The parameter γ in eq. (29) is used for compactness, and is defined as:

image (33)

This term also becomes singular for q3 = q4 = 0, which are retrograde equatorial orbits as explained above.

The variable p is the ratio between C and the velocity perpendicular to the radius vector image and ωi is a body-fixed angular velocity component:

image (34)

The angular velocity about the e2-axis, ω2, is zero, because there is no velocity component out of the instantaneous local orbital plane.

Orbital Stability and Monte-Carlo Analysis

Stability has many meanings. In the context of the current discussion we will look at the dynamical stability of satellite orbits, and in principle we would like to answer the following questions. What is the influence of small changes in initial state on future states? How does a small change in velocity affect the future states of the satellite? Is an Earth orbit stable and in what sense is it stable? A Monte-Carlo analysis may help us to answer these questions.

Small changes in initial values have small influence on a satellite’s orbit size and shape, i.e., a two-body orbit is orbital stable. However, small changes in initial values have large influences, increasing with time, on the position of a satellite in its orbit. The orbit is Lyapunov unstable (or dynamically stable) in the along-track direction, which can be shown as follows.

Imagine two satellites with unperturbed circular orbits around Earth. The orbits are almost equal, but have a small difference in radius, Δr. The angular velocity, n, of these satellites is given by:

image (35)

and thus the difference in angular velocity is:

image (36)

The relative angular position, Δθ, will thus increase with increasing time (Battin, 1999):

image (37)

This shows that no matter how small the difference in radial position, the difference in angular position along the orbit will always increase with time. In other words, the orbit is unstable in the along-track direction. This along-track dynamical instability leads to a numerical instability for orbit computations. For conventional numerical integration this causes an unknown error in the along-track direction.

These analytical expressions will, of course, not hold for perturbed orbits although they may give a good first estimate. For perturbed orbits, though, one numerically propagates the trajectory for the two different initial states. One can easily extend this concept to the uncertain measurement of the initial satellite position and velocity. Suppose the initial position and velocity are known with an error er and ev, respectively. This means that the initial state could be anywhere in the interval (r0-er,r0 + er) and (v0-ev,v0 + ev), where r0 and v0 are the actual initial position and velocity, respectively. Let us indicate this interval with (r) and (v).

To predict where the satellite will be after a certain period of time we need to take all possible initial conditions (r) and (v) into account. That means that if we assume a certain distribution of (r) and (v), e.g., a normal (= Gaussian) or uniform distribution, we could sample the respective intervals n times and do an orbit propagation for each of those samples. This is the basis of a Monte-Carlo analysis.

As an example we will show the results of a Monte-Carlo simulation, where 10,000 initial points have been randomly picked from a given 3-dimensional position interval with widths of 1 km. The resulting positions after 1.5 orbits are shown in Figure 8.3.8 (left). The red Monte-Carlo solution set is positioned along the blue nominal orbit; this shows the along-track instability. It grows from about 30 km after 1.5 orbits to over 400 km after 20.5 orbits. Note that the effect of position and velocity changes is largest after n + 0.5 orbits.

image

FIGURE 8.3.8 Monte-Carlo generated solution space for a LEO satellite, after 1.5 orbits, with an initial position width of 1 km in the x, y and z components (left), and an initial velocity width of 100 m/s in the x, y and z velocity components (right). (Römgens, 2011)

Figure 8.3.8 (right) shows a similar solution set, but now caused by a larger uncertainty in the velocity of 100 m/s, after 1.5 orbits. Again, the along-track instability is clear from the shape of the solution set that spreads out mostly in the direction along the nominal orbit. In fact, after 20.5 orbits the satellite can be located almost everywhere along its orbit. When the uncertainty in the velocity is 1 km/s in all components, the solution encloses almost the full orbit after only 1.5 revolutions.

Current collision-detection methods are based on estimates of the collision probability. As we have seen before, this collision probability depends on measurement errors in the state of the satellites at time t0. These errors are estimated by propagating the (assumed) known error probability density function at a certain starting time t0 to time t1, the estimated time of closest approach.

This probability propagation is traditionally done using analytical or conventional numerical integration methods. These methods introduce errors in the predicted trajectory at every integration step. These errors, however small they may be, can grow into much larger errors, mainly due to the along-track instability of satellite orbits. Furthermore, the satellite position is only predicted at fixed-points in time and not over a continuous time interval. Thus, they do not cover every moment that collisions may occur.

Given the limitations of Monte-Carlo analysis coupled with the fact of discrete-time prediction this may not be the best approach to try and detect possible collisions between satellites. Therefore, in the next section we will introduce an alternative method.

Verified Interval Propagation

Verified integration methods form a novel type of integration methods that can both enclose integration errors and integrate for a continuous range of (uncertain) initial values and parameters (Römgens et al., 2011). Moreover, they provide solutions valid over a continuous time interval rather than for discrete time points. Verified integration has been developed by a small group of mathematicians over the last 30 years, but has not yet been applied to many real-world problems (Neher, 2005).

Verified interval orbit propagation is the application of verified integration to solve the equations of motion that model satellite trajectories. This group of integration methods can integrate for a range (interval) of initial positions and velocities, and for a range of perturbing forces (parameters) that are applied during the integration time. It creates a guaranteed satellite trajectory enclosure, i.e., a 3-dimensional trajectory enclosing all possible trajectories corresponding to a range of initial conditions and forces. Moreover, it guarantees to include all numerical integration and rounding errors, thereby making the result verified, and all this with a single integration. An example of such an orbit enclosure is shown in Figure 8.3.9, where adjacent cuboids enclose the satellite’s position during different time intervals.

image

FIGURE 8.3.9 Example of an interval orbit enclosure created using verified interval propagation. (Römgens, 2011)

The next logical step in the development of collision-detection software is the realization that if the interval orbit enclosures of two satellites intersect, it may indicate that the satellites are at the same position at the same time. However, this thought may appear to be more complex than it seems.

Figure 8.3.10 shows the interval position enclosure and Monte-Carlo fixed-point solution set computed using MEE, for a circular equatorial LEO at 400 km, after 10 hours. The cuboid represents the interval enclosure, valid during a time interval of 1 s. The Monte-Carlo solution set, created using 100,000 individual conventional integrations randomly picked (uniform distribution) from within the interval initial values and parameters, lies within the interval enclosure.

image

FIGURE 8.3.10 Example of an interval solution enclosure, and a Monte-Carlo estimated true solution set plotted as the dot-cloud. (Römgens, 2011)

The figure also shows the overestimation that occurs when enclosing a non-convex 3-dimensional shape in a 3-dimensional interval. The Monte-Carlo solution set can be considered a good indication of the true solution set, although it does contain numerical integration errors. All other positions inside the solution enclosure are non-solutions. Although the true solution set will always be inside the enclosure, practical application of verified integration becomes useless when there are relatively many non-solutions inside the solution enclosure, e.g., knowing that a satellite is within a cuboid enclosing Earth and all LEO space around it is not very useful.

However, also the opposite is true: if the two enclosures do not intersect it is certain that the two satellites have not collided. And, since the true solution is mathematically guaranteed to lie inside the enclosure, the conclusion about not colliding is absolute.

In terms of orbital-state model, we have the three models as discussed previously in the section “Perturbed-Orbit Propagation.” Vittaldev et al. (2010, 2012) concluded that for perturbed orbits, the USM performs better than Cartesian coordinates for both fixed-step and variable-step integration, in some cases by more than an order of magnitude. Römgens et al. (2011) found that also in interval formulation the USM is superior to the Cartesian model. Moreover, in some cases the MEE perform even better. Therefore, a combination of USM and MEE is the best way to minimize the overestimation error mentioned earlier and to postpone solution explosion.

In Figure 8.3.11, the interval enclosures of different state models (MEE, USM, Cartesian and Kepler elements) are shown for an unperturbed orbit, as well as the intersection of the four models. As the theory of interval analysis guarantees that the actual satellite position lies within each individual enclosure, this means that the smallest possible enclosure will contain the satellite position. So if we use a hybrid state model we can use the intersection of the enclosures to reduce the overestimation. For the perturbed-orbit case we will use a combination of USM and MEE. In practise this means that we propagate the satellite orbit with both state models, determine the intersected enclosure and update the enclosures of the individual models.

image

FIGURE 8.3.11 Hybrid interval enclosure for an unperturbed orbit, composed of individual enclosures from different state models. (Römgens, 2011)

To apply hybrid verified interval propagation to satellite-collision detection we will try to reproduce the largest and most notable collision in Earth orbit, i.e., the collision between two intact satellites, the operational communication satellite, Iridium-33, and the inactive navigation satellite, Cosmos-2251. This collision occurred at 16:56 UTC on February 10, 2009, at 789 km above the Taymyr Peninsula in Siberia (Iannotta, 2009).

For the sake of the example, the uncertainty in initial conditions has been taken small: the initial interval widths used are 10 m in all position components and 0.1 m/s in the velocity components. Drag and the J2-effect are the perturbations that are explicitly modeled. The uncertainty in the density is enclosed in an interval with a width of 2 × 10−6 kg/m3, J2 is taken as the interval (1.08262 × 10−3, 1.08264 × 10−3), and an uncertainty in all other perturbations is modeled by an interval with a width of 2 × 10−9 km/s2. Both orbits are propagated using verified interval integration with an initial step-size of 10 s.

All interval trajectories are checked for intersections with other trajectories. Two possible collisions are found per revolution for the initial integration with a step-size of 10 s. One is at the location where the actual collision occurred and one is at the close approach area of both orbits at the opposite side of the globe (Figure 8.3.12).

image

FIGURE 8.3.12 Nominal orbits and collision location above North Pole (left), and the verified interval trajectories and detected interval collision between Cosmos-2251 and Iridium-33 (right). (Römgens, 2011)

When the step-size (time interval) is reduced to 1.25 s, possible collisions remain around the location of the actual collision, but all close approaches on the Southern Hemisphere are determined safe. The first possible collision is detected around 15:15 UTC, the second between 16:55:48.75 and 16:55:51.25. This second collision is the exact time of the real collision. Reducing the step-size to 0.3125 s removes the possible collision at 15:15. The real collision can still not be ruled out, which was expected since the satellites did actually collide.

The interval trajectories and the detected collision interval computed using a step-size of 5 s are shown in Figure 8.3.12. The solid and dashed lines indicate the nominal trajectory of Cosmos-2251 and Iridium-33, respectively. The time dimension of the interval trajectory is indicated by the opacity of the intervals, where a darker interval indicates a later moment in time. Projections of the interval trajectory and the black collision interval are shown on all three planes. This demonstrates that a true collision can successfully be detected by verified collision detection.

Conclusions

Based on the topics discussed in the previous sections, a number of top-level conclusions can be drawn. Real-time conjunction analysis is crucial to safeguard the operational lifetime of any user satellite. An efficient implementation of this conjunction analysis should include a two-step approach as discussed above: the first-order analysis, which only takes nominal orbital positions into account, and the detailed analysis, which also includes orbital uncertainties. As for the first step, the use of filters and sieves is highly recommended. For the detailed step, it is advised to use a combination of modified equinoctial elements and the unified state model, instead of Cartesian components. Verified interval propagation gives position and velocity enclosures that are valid over a certain time span, instead of at a discrete epoch. Finally, and most important, any space operator should clean up his/her satellite after completion of its mission, to continue the use of space by future generations.

Acknowledgements

The authors gratefully acknowledge the contribution of MSc. students Jonathan Leloux and Bart Römgens.

References

1. Altman SP. A unified state model of orbital trajectory and attitude dynamics. Celestial Mechanics and Dynamical Astronomy. 1972;6(4):425–446.

2. Battin RH. An Introduction to the Mathematics and Methods of Astrodynamics. AIAA Education Series 1999.

3. Betts JT, Erb SO. Optimal Low Thrust Trajectories to the Moon. Journal on Applied Dynamical Systems. 2003;2.

4. Cornelisse JW, Schöyer HFR, Wakker KF. Rocket Propulsion and Spaceflight Dynamics. Pitman 1979.

5. CSSI. SOCRATES, Satellite Orbital Conjunction Reports Assessing Threatening Encounters in Space. 2011; < http://celestrak.com/SOCRATES; 2011; > Consulted 13.02.12.

6. Hintz GR. Survey of Orbit Element Sets. Journal of Guidance Control and Dynamics. 2008;31(3):785.

7. Hoots FR, Roehrich RL. SPACETRACK Report No 3: Models for Propagation of NORAD Element Sets. Colorado Springs, CO: Technical Report, US Air Force Aerospace Defense Command; 1980.

8. Hoots FR, Crawford LL, Roehrich RL. An analytical method to determine future close approaches between satellites. Celestial Mechanics and Dynamical Astronomy. 1984;33:143–158.

9. Iannotta B. U.S Satellite Destroyed in Space Collision. 2009; < www.space.com/news/090211-satellite-collision.html; 2009; > Consulted Feb. 2009.

10. Kelso T. Celestrak. 2010; <http://celestrak.com/> Consulted 13.02.12.

11. Klinkrad H. Space Debris, Models and Risk Analysis. Springer 2006.

12. Klinkrad H. ESA Space Debris Office. 2009; <www.esa.int/esaMI/Space_Debris/SEMXP0WPXPF_0.html> Consulted 13.02.12.

13. Kuipers JB. Quaternions and rotation sequences. Princeton Univ. Press 1999.

14. Leloux J. Filtering techniques for orbital debris conjunction analysis. MSc thesis report Delft, The Netherlands: Delft University of Technology; 2012; <http://repository.tudelft.nl> Consulted 23.02.12.

15. Neher M. From interval analysis to taylor models – An overview. Proceedings IMACS (pp. 102–0658) 2005.

16. Römgens BA. Verified Interval Orbit Propagation with an application to Satellite Collision Detection. MSc. Thesis Delft University of Technology 2012; <http://repository.tudelft.nl> Consulted 23.02.12.

17. Römgens BA, Mooij E, Naeije MC. Verified Interval Orbit Propagation in Satellite Collision Avoidance. AIAA-11–6653 Portland, Oregon: From: AIAA Guidance, Navigation, and Control Conference; 2011.

18. USSTRATCOM. Space Track. 2012; <www.space-track.org/> Consulted 13.02.12.

19. Vallado DA, Crawford P, Hujsak R, Kelso T. Revisiting Spacetrack Report #3: Rev 1. In: Proceedings of the AIAA/AAS Astrodynamics Specialist Conference. 2006; Keystone, CO.

20. Vittaldev V, Mooij E, Naeije MC. Performance Aspects of Orbit Propagation using the Unified State Model. AIAA-10–7658 Toronto, Ontario: Proceedings of the AIAA/AAS Astrodynamics Specialist Conference; 2010.

21. Vittaldev V, Mooij E, Naeije MC. Unified State Model theory and application in Astrodynamics. Celestial Mechanics and Dynamical Astronomy 2012;. doi 10.1007/s10569-011-9396-5.

22. Walker MJH, Ireland B, Owens J. A set modified equinoctial orbit elements. Celestial Mechanics and Dynamical Astronomy. 1985;36(14):409–419.

23. Wertz JR, Everett DF, Puschell JJ. Space Mission Engineering: The New SMAD. Microcosm 2011.

8.4 Collision Avoidance Maneuvers for the International Space Station (ISS)

William Lark Howorth

Early History of Orbital Debris Avoidance

NASA has long known the potentially catastrophic consequences of any collision between a manned space vehicle like the Space Shuttle or Space Station and any piece of orbiting debris or other satellite large enough to be tracked. However, the agency’s stance toward addressing the risks from these close, high-speed “conjunctions” by maneuvering out of harm’s way have evolved greatly over the years. Until 1986, NASA’s manned spaceflight operations team and the Shuttle Program concluded that the cost to implement a regular collision avoidance process by performing a debris avoidance maneuver outweighed the benefit the program would realize by establishing this capability. The program did not have enough evidence at the time to show that establishing a process to avoid tracked orbital debris would be worth the effort and expense, given what were generally low risks for the short durations of Shuttle missions, and the then-incomplete understanding of how much such a process would be of benefit.

In subsequent years, NASA made substantial improvements in understanding the risks and benefits, and began establishing a method of avoiding such collisions. Cooperation between NASA and the United States Space Command (USSPACECOM, which later became U.S. Strategic Command, or USSTRATCOM), the organization charged with tracking and maintaining knowledge of all the tracked objects in Earth orbit, yielded a better understanding of how accurately orbital debris trajectories could be predicted. Following a gradual improvement in understanding, teams from USSPACECOM and NASA collaborated on agreements that defined the methods by which USSPACECOM would notify the Shuttle team for conjunctions up to 36 hours in advance of the potential collision.

These operational triggers initially were based on distance and geometry: when USSPACECOM predicted that any object in their catalog was to enter a fixed, box-shaped volume around the Shuttle, the vehicle flight dynamics team would then take action as directed by internal NASA “flight rules” to assess the threat and react appropriately. The size of this action box was generic, and based on large, worst-case debris uncertainties that were independent of the real uncertainty of the specific debris object in question. The protection volume used guaranteed that no object passing outside it could ever have more than a 1-in-100,000 chance of colliding with the Shuttle. This also meant that well-tracked and predictable debris objects that passed barely within this box could still have very small collision probabilities, even if the Shuttle performed no avoidance maneuver. In the end, a big, conservative action volume requires maneuvers to avoid objects that might actually be little threat; but since the Shuttle was a vehicle capable of moving out of the way at the last moment, commanded by a single-nation flight control team, these maneuvers were minimized in quantity, and the operational consequences were generally small.

However, by the early 1990s the impending arrival of the International Space Station required a new way of assessing and avoiding collisions with orbital debris. Several considerations forced a new approach to the problem:

• The ISS is on orbit year-round, for perhaps more than two decades, making it much more vulnerable to collision than the Shuttle simply due to exposure time.

• The ISS requires complex interaction among six national and continental space agencies to operate, and notably requires coordination with Russia, who controls the propulsion system of ISS.

• It is the largest satellite ever put into orbit, making it uniquely exposed to the ambient debris flux.

• The mass of ISS itself means that over 100 kg of costly propellant are required just to execute a 1 m/s avoidance maneuver.

• About a dozen governmental and commercial visiting vehicles are on the manifest to arrive and depart ISS every year, all of which desire a stable target for rendezvous.

With all these factors, the process of orbital debris avoidance for ISS could be a very costly enterprise if not done right. Therefore, NASA needed to find a way to reduce the number of avoidance maneuvers while keeping the laboratory and its six humans safe. Minimizing maneuvers while protecting this national and global asset became the urgent goal.

A New Way to Determine the Risk of Collision

In the early 1990s, as solutions were under development, it became clear that the one-size-fits-all volume-based “box” method used by the Shuttle to quantify the risk of collision from a specific object would not be adequate for ISS. NASA believed it could achieve a great benefit by adjusting the maneuver threshold based upon the specific prediction uncertainties of the conjuncting debris object and of ISS on the day in question.

This concept circled NASA back to the core principle that yielded the Shuttle avoidance box at the beginning: collision probability. Could a process be established for which the NASA operations team could compute the probability of collision for a single, specific conjunction? Research at the NASA Johnson Space Center showed that this was possible, as long as the predicted state-vectors and uncertainties for both the debris and ISS at the time of closest approach could be estimated. With certain assumptions that simplified the problem, this research showed that the collision probability could be calculated analytically using those state vectors and covariances as inputs. This new capability meant that a method of calculating the collision risk for a specific conjunction was indeed viable in an operational environment.

This all required the availability of orbital debris prediction uncertainties, a capability that didn’t exist in the early 1990s. These uncertainties, or covariances, are a result of tracking errors, along with unpredictable changes in space environmental conditions affecting an orbital trajectory. These factors combined yield an uncertainty ellipse both for the debris and for ISS at the time of closest approach. Managing the uncertainty was a crucial piece of the new probability-based method that NASA wanted to implement: for this reason, the agency’s analysts worked tightly with the U.S. Air Force to develop a method of computing covariances and verifying their accuracy. Once this capability was proven, NASA and USSPACECOM agreed on a special formalized notification to convey all the information required to compute the risk of collision: the Orbital Conjunction Message.

The ISS Debris Avoidance Operations Process

Knowing how to compute the probability of an object colliding with the Space Station is one critical step. But the key of a collision avoidance strategy is determining how to incorporate that risk assessment into an operational capability, or deciding exactly when to start considering evasive action, and what threshold to use to actually trigger it. The operations team can reduce the risk to the vehicle by lowering the threshold that triggers an avoidance maneuver. Conversely, the team can reduce the interruptions to operations by raising the maneuver threshold, a decision that increases the long-term risk to the asset. For the ISS, this sort of trade-off is part of a risk mitigation strategy that the operations community and program management have to accept (see Table 8.4.1).

Table 8.4.1

Establishing a strategy for ISS collision avoidance requires weighing safety and operational impacts

Image

The ISS program debated these kinds of issues, and finally selected the two maneuver thresholds that appear the best balance between safety and operational impact:

• “Red threshold”–1 × 10–4: For any conjunction for which the probability of collision exceeds 1-in-10,000, the operations team must move the ISS out of the path of any conjuncting debris, unless such an avoidance maneuver in itself would cause substantial risk to the vehicle or crew. Two examples of waiving a red threshold maneuver include the case when such an operation might lead to the failure of a crewed visiting vehicle to rendezvous with ISS, or when the ISS propulsion system is incorrectly configured for a maneuver.

• “Yellow threshold”–1 × 10–5: For a conjunction for which the probability of collision exceeds 1-in-100,000, the flight control team must maneuver ISS to safety unless it causes significant impacts to operations.

Over the history of the ISS program, approximately one avoidance maneuver per year has been performed. Studies performed by the ISS trajectory operations team showed that during a 5-year period (2003–2007), had the original Shuttle program box method been used for ISS, 22 debris avoidance maneuvers would have been required. The new probability-based technique on the other hand resulted in the execution of only one. The probabilistic risk assessment for conjunctions has undoubtedly shown a substantial decrease in disruption to ISS operations while maintaining safety.

Notification Criteria

The first step of the operational flow in collision avoidance operations is the initial screening of the Space Station against the other objects in the debris catalog performed by USSTRATCOM, and the subsequent notification of a conjunction to the Flight Control Team in Houston. The thresholds at which USSTRATCOM warns NASA of a possible threat are variously called “notification criteria” or the “alert box”, and the concepts and thresholds for this part of the operation have evolved over time as NASA’s ISS analysts adjusted to operational realities and learned from experience.

The NASA flight dynamics team that initially established the processes for coming ISS operations utilized Shuttle notification criteria as a starting point in initial trials from 1997–2001, using the Shuttle-Mir “Phase I” program as a test bed. However, with the various considerations that made ISS different from Shuttle, the operations teams from NASA and Russia continually modified the original Shuttle notification thresholds into what have become the current criteria for the “kickoff” of an ISS conjunction scenario:

• USSTRATCOM routinely performs a screening of ISS against all tracked objects in orbit 72 hours into the future. This 3-day horizon provides sufficient time for trend analysis after the initial identification, and gives enough warning for the ISS multilateral operations team to coordinate an avoidance maneuver, if required.

• After any of these 72-hour screenings, USSTRATCOM notifies NASA’s ISS trajectory operations team of any conjunction within a volume around ISS bounded by (±) 2 km in the radial direction (altitude); 25 km in the along-track, or tangential, direction; and 25 km in the cross-track direction. If a conjunction violates this box, NASA start receiving high frequency updated predictions on the debris via Orbital Conjunction Messages, and then engages in trend analysis and quality checks all the way through to the time of closest approach. During this time, NASA performs regular computations of the probability of collision with each update from USSTRATCOM.

• If the predictions of a conjunction indicate the debris object will pass within an even smaller volume, bounded by 750 meters of ISS in the radial direction, NASA’s trajectory operations team notifies the entire multilateral ISS operations community. Experience has shown that even at the low-altitude, high-drag environment in which ISS flies (350–420 km), 750 meters is a sufficiently large radial screening gate to capture virtually all objects that could threaten ISS within 3 days.

• Once the international community is notified of the risk, two tasks begin: the operations teams begin considering what aspects of crew, vehicle, and mission planning need to be adjusted in case a maneuver turns out to be required; and the trajectory teams begin determining exactly what size and direction of avoidance maneuver would safely avoid the threat.

In most cases, this process guarantees that any object that could collide with ISS is identified early enough for the operations team to take action, while minimizing the likelihood of false alarms. However, there are exceptional situations in which even these safeguards allow an object to enter the alert boxes late in the process and disrupt the normal flow of operations, perhaps even precluding the capability of moving the vehicle to safety. Causes for such “late notification conjunctions” include:

• Debris objects that are difficult to track because of their size, material composition, or orbit. Large gaps in available tracking data can make an object’s orbit harder to determine well.

• Debris objects that have trajectories that are difficult to predict because of an extremely low mass-to-area ratio that makes them very susceptible to small changes in atmospheric or solar activity. Atmospheric drag affects these objects up to 200 times as much as the ISS. For this reason, these objects have proportionally unpredictable orbits – occasionally so much that the conjunction notification gates are violated late in the process.

• A perturbation of ISS that unexpectedly pushes it into the path of a piece of orbital debris. This includes attitude maneuvers, docking or undocking activities, and off-nominal translational maneuvers, or reboosts. After every perturbation to the orbit of ISS, the operations teams perform a new screening of the vehicle against the debris catalog, and in some cases this yields a late-notice conjunction.

In these late-notification scenarios, there typically is still enough time to assess the conjunction and execute an avoidance maneuver, even though it might require a compressed timeline. There are scenarios, however, where the notification comes with so little time before the time of closest approach that the only option available is to safe the crew in their Soyuz emergency escape vehicles and prepared to depart, in case the debris strikes ISS. This scenario has well-scripted procedures for the crew, and, fortunately, the ISS team has only rarely had to exercise such a “shelter-in-place.”

Planning and Executing a Debris Avoidance Maneuver

Once the ISS team has received notification that a piece of debris will conjunct with ISS within 72 hours, the flight dynamics team at NASA begins the ongoing process of computing the probability of collision. The team will frame those results within the maneuver thresholds that the flight rules define, whether the results are “red”, “yellow”, or neither, and begin the process for planning a debris avoidance maneuver if required. For conjunctions that show little risk of collision, where the Pc is extremely low, usually very little activity among the flight controllers is required other than monitoring by the flight dynamics team. However, for higher-risk situations, planning for a maneuver and evaluating the probability of collision occur simultaneously throughout the scenario.

After learning from some early mistakes in the program that included trying to execute ground procedures too quickly, Russian and NASA operations teams evolved to a standard, minimum lead time to execute a debris avoidance maneuver, and documented it in the multilateral Operations Interface Procedures. Russia has authority over the propulsion system of ISS, so NASA has to issue a formal Planning Process Change Request 28.5 hours before the time of closest approach in order to safely plan the maneuver. Subsequently, NASA has to develop, verify, and uplink the commands to the vehicle. This sequence, if fully executed, ends with a debris avoidance maneuver about 135 minutes (one orbit and a half) before the conjunction.

The avoidance maneuver is evaluated all along the 28.5 hour window. If the trajectory teams believe the risk is high, the ISS operations teams will proceed toward an avoidance maneuver. Between this point and the planned time of ignition, NASA continues to evaluate the collision risk, giving an opportunity to cancel the operation if the risk drops convincingly below the maneuver thresholds defined in the flight rules. There have been several cases in the history of the ISS where a red threshold violation raised 28 hours before the time of closest approach vanished right before the conjunction due to the elimination of large prediction errors as the conjunction approaches. In general, the later the operations team can wait to make a final decision on whether to maneuver, the less likely the maneuver itself will be required. For this reason, and in order to reduce the impact of late-notification conjunctions, the ISS program partners plan to develop new vehicle capabilities that permit a much shorter timeline for planning and executing a debris avoidance maneuver.

When a debris avoidance maneuver is required, two major considerations go into determining exactly the delta velocity: (a) what is required to ensure the maneuver itself keeps the ISS safe from the debris; (b) whether the maneuver negatively affects downstream operations because of the unexpected change in the orbit of the ISS. In order to keep the ISS safe from the debris, trajectory specialists generally aim to perform an impulse between 0.5 and 1.0 m/s in the posigrade direction as often as possible. A velocity of 0.5 m/s is considered a minimum to guarantee overcoming any trajectory uncertainties of the debris and the ISS. For other satellites operating in a lower-drag environment at a higher altitude, 0.5 m/s is often a vastly larger impulse than is required to guarantee safety.

In addition to the safety consideration of the maneuver, ISS specialists must also consider the substantial impacts to the wide array of vehicles due to launch, rendezvous, dock, undock, and deorbit with the station up to two months after the debris avoidance maneuver has been executed. Soyuz and Progress, for example, have certain restrictions on rendezvous altitude and phase angle. Even if a particular vehicle does not have a defined constraint, changes in the orbit of the ISS impact their flight profiles, and perhaps even the date on which they can launch. That is, these vehicles have limits on where the ISS can be in its orbit in order to guarantee a rendezvous and docking in, say, two days, given the program-determined launch date for that vehicle. For a crewed vehicle like Soyuz meeting these constraints is extremely important, while for unmanned cargo vehicles there is typically some flexibility in their rendezvous duration before arriving at the ISS.

Keeping the ISS safe from collision is of utmost importance when designing the impulse of a debris avoidance maneuver, but the ISS trajectory teams must also take into account substantial secondary requirements in order to maximize the efficiency of the program as a whole.

Summary

The current methods the ISS team uses to assess the risk of colliding with tracked objects in space, and the methods for mitigating those risks, came out of many years of experience, going back to the early phases of the Space Shuttle program and are still evolving. Keeping the vehicle safe from collisions has demanded adapting these techniques to the changing conditions of the ISS program, and to learn from the experience accumulated over time. The addition of new international partners, the ever-increasing size of ISS and its changing propulsive capabilities, and the increasingly hazardous orbital debris environment means that close collaboration between operations, flight dynamics, program management, and military are critical to successfully reduce these risks. This is just as true going into the future, as the entire ISS partnership must continue improving methods of protecting ISS and its supporting vehicles from collision, as the population of space debris and satellites will increase in coming years.

8.5 Safe On-Orbit Manoeuvres Design

8.5.1 Rendezvous and Docking Operations

Wigbert Fehse

Introduction

Rendezvous and docking/berthing involves two spacecraft, chaser and target, a controlled approach to contact and the subsequent coupling between the two vehicles. Any contact outside the margins set for position, velocities and angular rates constitutes a safety critical collision. Collision safety risks, causes and consequences and the measures for protection are the main issues of this subchapter. In this context, external trajectory disturbances, navigation and thrust errors, the safety features and failure possibilities of the onboard systems and communication links, and the possibilities for the design of safe approach trajectories are discussed.

In the early days of space exploration, when the Soviet Union and the United States had launched their first satellites, both competing spacefaring powers realized that for the majority of space missions involving more than a single satellite, the capability of rendezvous and coupling of two space vehicles would be required. This technique is needed for:

• assembly in space to obtain larger units;

• joining of a vehicle in orbit by an ascending one to exchange crew and goods;

• rescue, repair, retrieval of spacecraft.

Projects such as the Moon-landing and manned orbital stations, pursued by both space powers since the early 1960s, right after launching the first humans into space, required the mastering of rendezvous and docking (RVD) in space.

In preparation of these complex mission goals, the United States conducted their “Gemini” program (Wikipedia, Project Gemini, 2013) and the Soviet Union their “Vostok” and later their “Soyuz” programs in the 1960s. As part of these programs, both sides pursued the development and demonstration of in-orbit rendezvous and docking (Siddiqi, 2000).

For a number of reasons, the development of the orbital rendezvous and docking techniques continued with the two space powers in somewhat different directions:

• The American space program selected manual control of the approach by an astronaut for at least the final approach prior to contact.

• The Russian space program selected automatic control of the approach by an onboard control system, with the possibility of monitoring and of manual steering by a human operator (in the spacecraft or on ground) as backup.

The manual approach had an initial advantage of lower development effort and the better flexibility to adapt the approach strategy to various operational requirements in Earth and lunar orbits. It had also the advantage that humans can better react to unforeseen situations.

The automatic approach had the initial advantage that it could be applied for the rendezvous and docking of both manned and unmanned spacecraft. In manned spacecraft, it has the advantage of leaving pilot free to monitor and supervise spacecraft system and operations. Also, eventually an automatic system is more suitable to routine RVD operations of manned and unmanned vehicles in a space station scenario.

• The first orbital RVD was performed by the United States on 16 March 1966, when Neil Armstrong and Dave Scott docked their Gemini VIII under manual control to an unmanned Agena target vehicle, a rocket upper stage specially equipped for this mission with a docking interface and visual target patterns.

• The first automatic RVD took place on 30 October 1967 between two Soviet vehicles: Cosmos 186 and Cosmos 188. Both vehicles were unmanned early Soyuz-type spacecraft, which were equipped with the IGLA rendezvous sensor system.

RVD operations were extensively used: in the “Apollo” Moon-landing program (Wikipedia, Apollo program, 2013) by the US; and, after losing the race to the Moon, the Soviet space program used and improved the techniques and technology of automated RVD for operations in the “Salyut” Orbital Station program (Wikipedia, Soviet space program, 2013). During that time, on both sides not much was published concerning the technical details of the RVD systems, because of the strong political competition and because all information on guidance, navigation and control (GNC) systems were considered sensitive military information. Both space powers have continued until today with their initial choice of either manual or automatic control for the final RVD phase. The USA performed a large number of RV missions in their Apollo, Skylab and STS program in different orbits, with various types of targets, and different coupling techniques, including docking and capture and berthing of the target by a manipulator (see “Safety Risks during Rendezvous, Capture and Departure”). The Soviet/Russian space program on the other hand developed within its Salyut and Mir Space Station program the automated rendezvous and docking from experimental state to a routine operation.

With the exception of the Apollo–Soyuz Test Project (ASTP) (Wikipedia, Apollo–Soyuz Test Project, 2013), flown in 1975, up to the advent of the International Space Station program (ISS) (Wikipedia, International Space Station 2013) in the early 1990s, in both the American and Russian scenarios, all operations of chaser and target vehicles during rendezvous were under control of that scenario’s own systems, pilots and control centers. This simplified the issue of operational safety to a certain extent; and since all elements of the rendezvous and coupling systems and operations were developed by and under control of the same space power, one could rely on experience and rules of one’s own program and culture.

Whereas in a test program with two different partners, such as in the ASTP, one could make ad hoc agreements on all issues of joint operations, in an operational program such as the ISS, where many different space powers are involved and where in many of the rendezvous and coupling operations chaser and target belong to different sides, firm rules for the safety of operations needed to be established. Safety and mission success were in the early days of space operations not considered to be essentially different issues. However, in a mission without rendezvous and coupling with another spacecraft, mission success would be the highest value to be safeguarded. In a rendezvous mission with another manned spacecraft, which is owned by other partners, the safety of the partners (or better of their crew and vehicle) is of higher value than the mission success of the own vehicle.

The first rendezvous operations within the scope of the ISS program that included more than one side to control the process were the flights of the American Space Shuttle to the Russian Mir Station in the 1990s. They were performed in preparation of the ISS, to obtain more experience in all space station operations. These flights required development and implementation of joint operations during the approach and the use of the Russian docking mechanism by the US Orbiter. In addition, planning of rendezvous and coupling of vehicles of other space powers to the ISS, i.e. Europe and Japan, started at that time. With the experience of the Shuttle–Mir rendezvous operations and the forthcoming rendezvous of new vehicles in mind, a set of safety rules were developed (see “Safety Requirements for Rendezvous Missions” below) within the ISS program. This included the establishment of control zones around the ISS, as shown in Figure 8.5.1.1, where an outer “approach ellipsoid”, an inner “keep-out zone” and “approach corridors” are defined.

image

FIGURE 8.5.1.1 Control zones of the ISS. (Fehse, 2003)

The approach ellipsoid has an extension of ±2000 m in orbit direction (x-, V-bar direction) and ±1000 m in the other directions. The rule is that prior to any maneuver that will enter the approach ellipsoid, overall control must be handed over to the ISS control center.

The keep-out zone is a sphere with a radius of 200 m, which can be entered only inside the approach corridors. Prior to entering the approach corridor, the ISS control center must give its “go ahead”. In the case of docking, the approach corridor originates in the docking port; in the case of capture and berthing of the approaching vehicle by the station’s manipulator, it originates in the so-called “berthing box” (see Fehse, 2003, pp. 116/121). Depending on the reception range of the docking port, the corridor can have a half-cone angle of 15 deg at 200 m and 10 deg or less in the short range prior to docking. At any violation of the boundaries of the corridor, the incoming vehicle has to initiate a Collision Avoidance Maneuver (CAM); see “Safety Design of Onboard Control System” (p. 497) and “Active Collision Protection” (p. 508).

The first vehicles to follow strictly the new ISS safety rules (the US Space Shuttle and the Russian Soyuz and Progress vehicles continued basically with their proven RVD schemes) were the European “Automated Transfer Vehicle” (ATV) (Cornier et al. 1999), which on 3 April 2008 for the first time performed automatic rendezvous and docking to the ISS, and the Japanese “H2 Transfer Vehicle” (HTV) (Kawasaki et al. 2000), which performed automatic rendezvous with the ISS and was captured and berthed by the ISS robotic arm for the first time on 18 September 2009. Regular flight of these vehicles are planned to re-supply the ISS. At the time of writing, a second HTV has flown to and been berthed at the ISS on 27 January 2011 and a second ATV has flown and docked to the ISS on 24 February 2011.

Although rendezvous and docking operations have been performed so far only in low Earth orbit (LEO) and in a lunar orbit by the Apollo program, there is a strong interest in a number of other applications, which range from rendezvous in geostationary Earth orbit (GEO) up to new missions to the Moon and missions to Mars. For such missions, in particular when unmanned spacecraft are involved, automatic and even autonomous rendezvous and coupling techniques will be required. A large number of study programs, development programs and also in-orbit demonstration programs have been performed for the development of automated rendezvous and docking (some times even called “autonomous RVD”, although the level of autonomy was limited). In-orbit demonstrations of RVD techniques and technology included the Japanese “ETS-VII” mission (Kawano et al., 1998), the “Demonstration of Autonomous Rendezvous Technology” (DART) (NASA NESC 2006), and the “Orbital Express” (Mulder, 2008) missions performed by the US.

Whereas in manned missions safety of human life is the most important issue, safety requirements for unmanned rendezvous missions will depend on other values to be protected (see “Safety Requirements for Rendezvous Missions” below). Such requirements will have to be developed along with all onboard and ground systems and operations.

Rendezvous in GEO may become soon commercially interesting for the communication satellite market, to extend the operational life of satellites and to remove expired/failed satellites from a valuable position on the geostationary rim (Wingo, 2004; Kaiser et al. 2008). For safety considerations, servicing of communication satellites is a particularly interesting case. In contrast to the space station scenario in LEO, where safety of human life is the most important issue, with communication satellites in GEO the capital investment in the satellite and the daily revenue, which can be endangered by rendezvous and coupling operations, are the most important issues for the commercial enterprises operating the satellites (see “Safety Requirements for Rendezvous Missions” below). For this reason, the orbital disturbances in LEO and the disturbances in GEO are discussed in this chapter as potential causes of collision danger (see “Effects of External Disturbances on the Trajectory Evolution”, p. 487 ff.).

Safety/Security Issues during Rendezvous, Coupling and Departure Operations

Safety Requirements for Rendezvous Missions

Safety requirements focus on protecting the highest value aspects in the type of mission concerned. For the rendezvous, capture and departure part of a mission it is critical to determine how these high value elements may potentially be compromised during those operations and how the general safety requirements translate into specific technical requirements for

• approach/departure trajectories; and

• rendezvous control system design.

In manned mission scenarios (at least one of the two vehicles is manned), the highest value to be protected is the life of the crew in orbit and the possibility of their safe return to Earth. Rendezvous, coupling and departure operations must comply with the failure tolerance requirements, which for manned missions typically are, as established in the ISS program (NASA 1998):

1. No single failure shall lead to a loss of mission.

2. No combination of two failures shall lead to loss of life or to an disabling injury of crew and the loss of manned vehicles or major elements.

In this formulation of failure tolerance requirements, errors by human operators that will have repercussions on the ways of interaction by ground and onboard operators with the rendezvous and departure operations between chaser and target vehicles are also included. The above requirements can be translated for onboard systems and maneuvers into the more familiar fail operational–fail safe requirement. The resulting requirements for approach, capture and departure operations can be interpreted for the manned mission scenario as follows:

• Rendezvous and approach up to contact
After any first failure the rendezvous control systems on chaser and target must be able to continue/resume the mission and the trajectory must be safe (i.e. collision-free).
After a second failure, in any combination with a first failure, the target must be safe and

– the manned chaser must be able to return safely to ground.

– the unmanned chaser must be still have the necessary capabilities to be removed from orbit.

• Capture
After any first failure during the capture operations, it must possible to continue the capture process or to repeat the capture operations.
After the second failure, in any combination with a first failure, it must be possible to release the chaser from the target docking port and to remove it from the target. It is essential that the docking port of the target maintains the capability of receiving other approaching vehicles.

• Undocking
After any first failure during the undocking operations, i.e. opening of structural latches, opening of capture latches, separation, it must be possible to continue the operation. After the second failure, it must be possible to separate chaser and target and free the target docking port.

• Departure
After the first failure, the departing vehicle must be able to continue the mission and the trajectory must be safe.
After a second failure, in any combination with a first failure, the target must be safe and

– the manned chaser must be able to return safely to ground.

– the unmanned chaser must be still have the necessary capabilities to be removed from orbit.

It has to be noted that failure tolerance requirements are not necessarily additive for each of the four phases discussed above. A first failure in the rendezvous control system, e.g. during the approach phase, would also be considered a first failure in the departure phase.

For unmanned missions (both vehicles are unmanned), the highest values to be protected will depend on the type of mission, as shown in the following examples. The highest value can be:

1. The investment in both the target vehicle and the chaser vehicle, e.g. in case of a servicing mission in GEO, the still operational communication satellite, for which a significant investment has been made and which still produces revenue, and the servicer, for which a significant new investment in spacecraft and launch has been made.

2. The revenue, which is envisaged to be obtained for a particular time from the assets in orbit. This is a value particularly important for commercial satellites, e.g. communication satellites in GEO.

3. The investment in the total mission, e.g. in case of an assembly mission or a sample return mission to a celestial body.

4. The investment in the chaser only: this could be the case in a rescue/servicing mission to an incapacitated spacecraft. If the chaser in such case has additional targets to serve, it could be assumed that the incapacitated target is is of lower value.

In Case 1 the failure tolerance requirements resulting for the rendezvous, capture and departure operations would be similar to the case of an unmanned chaser in a manned mission scenario, i.e. after the first failure the rendezvous control systems of the chaser must be able to continue/resume the mission and the trajectory must be safe.

After a second failure, in any combination with a first failure, the target must be safe and the chaser must still have the capabilities to be removed from orbit.

In Case 3 the requirement of mission continuation after failure would be extremely important for mission success. This could lead, if affordable, even to a double failure tolerance requirement for mission success. However, a requirement for a keeping the target safe and removing the chaser from the orbit after a second failure, would in this case not make any sense, as the mission would end anyway, if rendezvous and capture cannot be performed any longer.

In Case 4, where it is assumed that the chaser can be used also for the servicing of other targets, the situation is similar to Case 2, with the extension, however, that after a second failure the chaser must still have the capabilities to be removed from the orbit.

These examples show that safety and failure tolerance requirements for unmanned missions cannot be defined once for all, but depend on the mission objectives and will have to be defined for each mission accordingly.

Safety Risks during Rendezvous, Capture and Departure

As addressed in the previous section, there are some important safety risks in a rendezvous and coupling operation. The potential consequences of these risks in manned and unmanned mission scenarios are stated hereafter.

1. Inability to complete the approach
This is a risk mainly for mission success. Inability to complete the approach could become a safety issue only if the incoming vehicle or its payload is immediately needed for further functioning of the target station or for rescue of crew. Otherwise, after a failure to complete the mission by a first chaser vehicle, a following vehicle could be launched to complete the task.

2. Collision between chaser and target
Rendezvous and docking comprises a complex sequence of operations that eventually leads to a controlled collision between chaser and target. Controlled are the position of the points of contact on both vehicles, the direction and amount of contact velocity and the residual angular rates. These parameters have to be controlled within the tight margins, which are necessary to achieve successful capture of/by the mutual interfaces on both vehicles. Any excess of the defined margins can lead to collision under conditions not designed for, with the risk of damage to the vehicles.
In manned missions, collision poses highest risk to the life of crew and for the integrity of chaser and target station. In unmanned rendezvous scenarios, it is the highest risk for the investment made in target and chaser vehicles, as discussed in the previous section.

3. Inability to capture
Inability to capture is to the first degree a risk to mission success. In manned missions, however, it can also become a safety risk for the crew on the target if the inability to capture prevents the delivery of resources urgently needed or potentially prevents the return of crew to Earth.

In rendezvous missions, capture can be achieved in two ways:

(a) The capture interfaces of the chaser will be driven by the Guidance, Navigation and Control (GNC)– and reaction control systems of the chaser into the capture interfaces of the target. After capture, the coupling interfaces, which are at the same location, will perform structural connection. This method is called “docking”.

(b) A passive capture interface on one of the vehicles is grappled by the end-effector of a manipulator on the other vehicle. The manipulator will then move the coupling interface of one vehicle into that of the other one so that structural connection can be achieved. This method is called “berthing”.


Structural coupling is the final part of the mating process and is, as all other operations, susceptible to failures. Structural coupling in space is, however, not different from coupling on ground and, therefore, does not need to be discussed in detail here. The risks are similar to inability to capture and inability to undock.
The inability to capture may be caused by either chaser or target. If it is due to a malfunction on the side of the chaser, the mission may be lost; however, a next chaser would be able to dock. If the malfunction is on the target side, the repercussions would be serious if the problem could not be solved. In such a case, the port could not be used any longer.

4. Inability to undock
For manned vehicles, inability to undock is a serious safety risk, as it would prevent crew from returning back to ground. Another risk is the blocking of a port on the target station, which would prevent further coupling of vehicles for crew exchange and replacement of consumables and would endanger the future functioning of an orbital station. This risk is valid for both manned and unmanned vehicles coupled to a manned target station.
In a fully unmanned mission scenario, inability to undock would be a risk for mission success in missions where undocking is part of the planned operations. There may, however, be mission cases where after initial mating undocking is not required. This would be the case, for example, in assembly and sample return missions.

5. Inability to depart
As under point 3, for manned vehicles, this is a serious safety risk, as it would prevent crew from returning back to ground. Another risk existing both in the manned and unmanned scenario is the risk of collision with the target, when a vehicle in close vicinity of the target has lost the capability to depart.

Safety requirements and safety risks of capture, coupling and release have been briefly discussed here for completeness. The present section will concentrate, however, on the first two issues. Capture and undocking are mechanical issues and, therefore, out of scope of this analysis; and the safety problems of departure are in essence not different from those of the approach.

Causes of Collision Danger

As discussed in the previous section, in a rendezvous mission collision is potentially the highest danger to the safety of life of crew and to the integrity of the two vehicles involved. Collision can occur when one of the two vehicles moves on a non-nominal trajectory. Since the target is mostly passive and does not perform maneuvers during rendezvous, trajectory deviations of the chaser pose the highest risk of collision danger. Non-nominal trajectories can be caused by:

• orbital disturbance,

• navigation error,

• thrust error,

• thruster failures,

• malfunction of the rendezvous control system or of one of its parts.

The global effects of these disturbances, errors and failures are described hereafter.

Orbital Disturbances

Drag of the residual atmosphere is in LEO, and solar pressure is in GEO, the largest orbital disturbance of rendezvous trajectories (Figures 8.5.1.4 and 8.5.1.5). Other effects, such as deviation of the gravitational field of the Earth from an ideal sphere play a role in the long range rendezvous operations, but do not create a safety danger.

Drag force of the residual atmosphere

The drag force acting on the body of a spacecraft is:

image (1)

CD = drag coefficient

A = cross section of body

where image is the orbital velocity, with ω is the orbital rate and r the radius of the orbit.

The acceleration of the spacecraft due to the drag force is:

image (2)

where m is the mass of the spacecraft. The actual value of the drag force image depends on the density of the residual atmosphere, which is changing with altitude and Sun activity, as shown in Figure 8.5.1.2.

image

FIGURE 8.5.1.2 Density of the atmosphere. (redrawn from ESA web-site 2010 / Doornbos 2009)

From Figure 8.5.1.2 it can also be seen that the density will vary to a certain extent along a circular orbit, due to the different temperatures in the day and night parts of the orbit.

Solar Pressure

In contrast to the drag of the residual atmosphere in LEO, which is always opposite to the flight direction (Figure 8.5.1.3), the direction of the solar pressure disturbance rotates with respect to the flight direction by 360 deg per revolution (in GEO per day) for the in-plane component and by ±23.5 deg per year for the out-of-plane component (Figures 8.5.1.4 and 8.5.1.5).

image

FIGURE 8.5.1.3 Drag, predominant in LEO. (Fehse, 2006)

image

FIGURE 8.5.1.4 Solar pressure, predominant in GEO. (Fehse, 2006)

image

FIGURE 8.5.1.5 Sun angle relations in GEO. (Fehse, 2006)

The force per unit of mass of the satellite produced by the solar pressure is

image (3)

where p is the radiation momentum flux, A is the cross section of the satellite normal to the Sun direction, m is the mass of the satellite and image is the Sun–satellite direction unity vector.

The radiation momentum flux varies periodically with the orbit of the Earth around the Sun and is

image

image

The x- and z-components of the solar pressure are:

image (4)

image (5)

where image, ω is the orbital rate and the time t starts at midnight.

The y- (out-of-plane) component is

image (6)

where β has a maximum of image at the solstices and becomes zero at the equinoxes.

Differential accelerations acting between chaser and target

In rendezvous operations, we are interested in the relative trajectory development, i.e. in the relative position and velocities between a chaser and a target spacecraft. Depending on the area to mass ratio of the spacecraft, it will be more or less accelerated by the disturbance force. If this ratio is the same for both vehicles, they will be affected in the same way and no effects will be noticeable in the relative navigation.

To describe the air-drag sensitivity of a body by a simple parameter for military applications on ground, the concept of a ballistic coefficient has been in use for a long time. This can also be used with the drag in LEO rendezvous applications and the “ballistic coefficient” image can be defined as:

image (7)

In the same way as for differential drag in LEO, for differential solar pressure effects in GEO we can also call the quotient of mass to area image, the “ballistic coefficient”. Since there is no flow of gas molecules, a drag coefficient image is not applicable for solar pressure.

Collision danger resulting from differential accelerations, caused by drag or solar pressure, can occur both during approach and departure.

Differential acceleration due to residual drag in LEO:

The differential acceleration between the chaser (index ‘c’) and the target (index ‘t’) will be according to eqs. (1) and (2):

image (8)

With the definition of the ballistic coefficient in eq. (7) image for air drag in LEO this equation becomes

image (9)

Differential acceleration due to solar pressure in GEO:

With the definition of the ballistic coefficient image, eq. (3) for the solar pressure becomes

image (10)

The disturbance acceleration noticeable in the relative trajectories will be:

image (11)

As mentioned above, if the ballistic coefficient of chaser and target vehicle are the same, the relative trajectories between the two vehicles will not be affected by the disturbance of the air drag or the solar pressure, whatever the absolute displacements with respect to the undisturbed case may be.

Calculation of Trajectory Deviations

In a circular orbit, the relative motion between two bodies, e.g. chaser and target, are described (in a frame centred in, and moving with, an undisturbed target point in orbit) by the well known Hill equations:

image (12)

where ω is the orbital rate and image is the mass of the chaser vehicle. The disturbance accelerations are image. This system of second order differential equations can be solved generally only by numerical integration.

For step changes of the velocity at start and for constant forces, the disturbed trajectory can be calculated by the Clohessy–Wiltshire (CW) equations, which are a particular solution of the Hill equations.

image (13)

The CW equations give acceptable results as long as the altitude difference between the two vehicles is image, the orbit radius. In orbit- (x)-direction, the curvature of the orbit will eventually lead to deviations in the navigation. Also, they provide sufficiently accurate results to describe the trajectory deviations caused by external disturbances, only as long as the disturbance forces/accelerations are constant over the time considered. For disturbance forces that are variable over the time considered, the Hill equations need to be integrated numerically.

Disturbances can be produced by external forces, such as residual air drag or solar pressure, and by forces produced by the spacecraft itself, such as thrust deviations and thruster failures. Trajectory deviations can further be caused by navigation errors, i.e. errors in the measurement of position and velocities. For the detailed equations applicable for each disturbance and error case the reader is referred to the reference by Fehse (2003). In the following only the results for some particular cases will be discussed.

Effects of external disturbances on the trajectory evolution

The effects of the drag force in LEO are easy to understand, as it acts always opposite to the orbit direction (Figure 8.5.1.3). As a result of the drag, the orbit of each real body with a cross section A will decay. Figure 8.5.1.6 shows for a 400 km orbit relative to the vehicle with the higher ballistic coefficient the trajectory evolution over three orbits of the vehicle with the lower ballistic coefficient for different image ratios (after release from active station keeping).

image

FIGURE 8.5.1.6 Example: motion with differential drag, (start: image, image). (Fehse, 2003)

The vehicle with the lower ballistic coefficient (lower mass to area ratio) will always decay faster. The lowest trajectory in Figure 8.5.1.6 shows the motion of a target, the ballistic coefficient of which is six times lower than that of the chaser. (Note: In the rendezvous community the x-direction is called “V-bar”, after the velocity vector, and the z-direction is called R-bar, after the radius vector).

As a result of the differential drag force, the shape of approach trajectories will be distorted and over long time their direction of motion can be reversed. An example is shown in Figure 8.5.1.7 for a tangential impulse of –0.06 m/s in a 400 km orbit with a ratio of 5 between the ballistic coefficients of chaser and target. The maneuver is intended to initiate a motion in x-direction away from the target, which is located at the origin of the coordinate system. As a result of the differential drag, the motion of the chaser reverses and comes after a couple of orbital revolutions back towards the position of the target. This trajectory evolution is safety critical, since an original maneuver, intended to lead away from the target, can without further correction eventually lead to collision.

image

FIGURE 8.5.1.7 Tangential impulse −0.06 m/s with differential drag image. (Fehse, 2003)

The effects of the solar pressure in GEO are much more complex than the effects of drag in LEO, since the direction of the pressure depends on the time of the day for the position of the satellite. A few examples of the trajectory evolutions for different starting times are related to an undisturbed point in orbit, shown in Figures 8.5.1.88.5.1.13. These figures are obtained by numerical integration of the Hill equations, since the CW equations are not valid, as explained above, for forces that are variable over time.

image

FIGURE 8.5.1.8 Release from SK in GEO at 00:00 h.

image

FIGURE 8.5.1.9 Release from SK in GEO at 06:00 h.

image

FIGURE 8.5.1.10 +Vx-boost relative trajectory at 00:00 h.

image

FIGURE 8.5.1.11 −Vx-boost relative trajectory at 00:00 h.

image

FIGURE 8.5.1.12 −Vx-boost relative trajectory at 06:00 h.

image

FIGURE 8.5.1.13 +Vx-boost relative trajectory at 06:00 h.

Figures 8.5.1.8 and 8.5.1.9 show examples of the trajectory evolution relative to the undisturbed case of a body disturbed by solar pressure. Three revolutions are shown and start times of 0 h and 6 h local time. An average value of image has been used for the solar pressure and a ratio of image for the ballistic coefficient. The start condition of the trajectory can be understood, as if the body was kept in forced station keeping (SK) at the undisturbed position until its release at start. The comparison of two figures shows the strong dependency of the trajectory evolution from the start time. The trajectory evolution will actually be different for any point in time, but will symmetrical to both sides of the 12:00 h point.

In Figures 8.5.1.108.5.1.13, examples for the relative trajectory evolution between a chaser and a target are shown after an initial tangential boost of ±0.002 m/s. Start times are again 0 h and 6 h local time. Realistic ballistic coefficients have been assumed for the chaser (servicer) of image and for the target (communication satellite) of image. The evolution over three orbits shows that only for less then half a revolution the trajectory shows some similarity with the undisturbed case, thereafter the influence of the solar pressure disturbance becomes predominant.

Although the absolute value of the solar pressure force in GEO is much smaller than that of the drag force in LEO, the effects are even more critical. This is for one part due to the fact that the orbital rate ω is by a factor of about 16 smaller than in LEO, and for the other part that the direction of the disturbance force is rotating by 360 deg per orbit. In any case, it can be concluded that in GEO open loop trajectories without several mid-course corrections or closed loop control of trajectory boundaries will be unsafe at ranges lower than a few thousands of meters.

Effects of navigation errors on the trajectory evolution

In orbit, the development of the trajectory of a body after the application of an initial velocity or force will be different from that on ground. On ground, we are used to linear propagation with the present velocities from a present position to a future one. In orbit, propagation has to be done according to orbital mechanics. This often has results that are contrary to our experience on ground. For example, an increase of velocity in forward direction will in orbit lead after short time to a motion in backward direction. In the same way, navigation errors can in orbit have different results than on ground. This in itself can be a source of mistakes, which might lead to safety critical conditions.

In an impulsive transfer to a different position, the error ellipsoid around the intended position will depend, in addition to the orbital disturbances discussed above, to its major extent on the navigation errors at start. The parameters needed for navigation in orbit are position, linear velocities, attitude angles and angular rates. Attitude and angular rate measurement errors, however, are not discussed here, as they do not lead to trajectory deviations.

In comparison, thrust errors (see subsequent section) will have much smaller effects than navigation errors.

For an assessment of the potential trajectory deviations, let us assume that the range to target is of the order of 10 km and that chaser has the following navigation errors:

• position error in x-direction image

• position error in z-direction image

• velocity error in x-direction image

• velocity error in x-direction image

• position error in y-direction image

It can further be assumed that the major part of the position error is due to bias and that the noise part is at least one order of magnitude lower. In many cases velocities and angular rates cannot be measured directly and have to be derived from the position measurements by averaging position measurements over time. Therefore, noise must be below a certain level.

In a 400 km altitude LEO orbit, the error evolution over one orbital revolution would be for position and velocity errors (without orbital disturbances) as shown in Figures 8.5.1.148.5.1.17. The chaser spacecraft is assumed to be on the target orbit and to start a free motion with the above errors. The nominal x- and z-position at start is set to zero, and trajectories are shown relative to the nominal start position (origin of frame). Results are shown for positive error values, for negative error values the curves are mirrored on both the x- and z-axes.

• Position errors in y-direction (out-of-plane) will evolve along the orbit with image (no figure), i.e. after half an orbit the trajectory deviation will be image.

• The position error caused by an initial x-measurement error of image does not result in any further error evolution over time. It remains always at the 100m x-distance on the target orbit altitude.

• The evolution of trajectory deviations, caused by measurement errors of the z-position, depends on the actual velocity in x-direction at start. Figures 8.5.1.14 and 8.5.1.15 show two particular situations, for which the initial position error is image.

– In Figure 8.5.1.14, the velocity in x-direction is that of a circular orbit 100 m below the target orbit, which yields after one orbital revolution a change in x-direction of image, i.e. about 950 m.

– In Figure 8.5.1.15, the z-position error is 100 m, but x-velocity is that of the target orbit, which yields after half a revolution about 1885 m in x-direction and 700 m in z-direction, and after one revolution about 3770 m in x-direction, returning to the initial 100 m in z-direction. The velocity difference between the two cases is about 0.17 m/s, which is of the order of magnitude of the assumed velocity measurement errors. The actual error case could be the combination of the measurement accuracy in z-direction with any velocity in x-direction within the limits of the x-velocity measurement accuracy.

• The evolution of trajectory deviations caused by x-velocity measurement errors is shown in Figure 8.5.1.16. At half an orbital revolution the trajectory has reached about 850 m in x-direction and 350 m in z-direction, and after one revolution the trajectory comes back to the x-axis (V-bar) at about 1700 m in x-direction.

• The evolution of trajectory deviations caused by z-velocity measurement errors is shown in Figure 8.5.1.17. In contrast to the x-velocity errors, the trajectory deviations due to z-velocity errors have a maximum in x-direction after half an orbital revolution and in z-direction after a quarter and three-quarters of a revolution. After one revolution the trajectory returns to the starting point.

image

FIGURE 8.5.1.14 Trajectory evolution due to position error image, with motion on parallel orbit.

image

FIGURE 8.5.1.15 Trajectory evolution due to position error image, with same velocity as target orbit.

image

FIGURE 8.5.1.16 Trajectory evolution due to velocity error image.

image

FIGURE 8.5.1.17 Trajectory evolution due to velocity error image.

In GEO, trajectory deviations due to velocity measurement errors are larger by a factor of 15–16 than in LEO, because of the fact that the orbital rate ω is 15–16 times lower than in LEO (depending on LEO altitude).

The figures show that measurement errors of the z-position and of the velocity in x-direction lead to the largest trajectory deviations. Typical approach strategies contain impulsive maneuvers of half an orbit duration. Figures 8.5.1.148.5.1.17 show that the trajectory errors with the above initial navigation errors could lead to trajectory deviations of nearly 1900 m in x-direction and 700 m in z-direction. It is obvious that with such deviations an approach by impulsive transfers to a range to the target closer than a few kilometers would be unsafe. As a rule of thumb, position measurement accuracy must be less 1% of the range to the target. Continuously increasing measurement accuracy with decreasing range will allow for intermediate corrections with improved accuracy. From a certain range downwards (in LEO a few hundred, in GEO a few thousands of meters) closed loop control of trajectories will become necessary to achieve a safe approach to the mating position.

Effects of thrust errors on the trajectory evolution

Compared to the navigation errors, the effect of thrust errors on the safety of the approach is relatively uncritical. Under a thrust error we understand here the deviations from the nominal thrust level and duration. Thruster failures, i.e. “thruster-open” and “thruster-closed” failures, are a different issue, and are discussed under “Thruster failures” below.

Thrust errors have a proportional effect on the trajectory, i.e. a 1% lower thrust will lead to a 1% shorter distances from the start position in x- and z-directions. For open loop impulsive maneuvers the thrust error sources that play a role can be reduced as follows:

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset