THE HIPPARCOS MISSION FLOOR VAN LEEUWEN Royal Greenwich Observatory, Madingley Road, Cambridge, UK E-mail address: [email protected] (Received 22 October, 1997) Abstract. A review is presented of the European Space Agency’s astrometric satellite project Hipparcos, for which the final data catalogues were published in June 1997. The emphasis is on those aspects that have or may have influenced the data as presented in the catalogues. It presents a brief review of the satellite, the aims of the mission with their relation to ground-based astrometry, and the mission history. This is followed by a description of the input data streams and a summary of the data reductions. Data files presented in the catalogues are described in the context of the data reductions, and explained in content and usage. The mission results comprise not only outstanding astrometric data on both single and double stars, but also an unique all-sky photometric survey which has been used for variability investigations. This has led to the discovery of thousands of variable stars. The Hipparcos mission was complemented by the Tycho experiment, providing a complete all-sky survey of astrometric and photometric parameters for one million stars down to magnitude 11, though with lower accuracies than obtained for the main mission. Astrometric and photometric data for a selection of 48 minor planets, the Jovian moon Europa and the Saturnian moons Titan and Iapetus were also obtained. The data quality verifications are reviewed and guidelines to the proper use of the Hipparcos data are provided, followed by some of the first scientific results of the mission. Experiences gained with this first ever space astrometry mission are considered in relation to a follow up mission for Hipparcos. Parallaxes measure stellar distances, other methods rely on assumptions. Key words: Data reduction techniques, Double Stars, Hipparcos, Satellite dynamics, Space Astrometry, Tycho, Variable Stars Abbreviations: ABM – Apogee Boost Motor; CCCM – Coil Current Calibration Matrix; CERGA – Centre d’Etudes et de Recherches Géodynamiques et Astronomiques; DMSA – Double and Multiple Systems Annex; ESOC – European Space Operations Centre; FAST – Fundamental Astronomy by Space Techniques; FFOV – Following Field of View; GCR – Great-Circle Reduction; IDT – Image Dissector Tube; IFOV – Instantaneous Field of View; ITF – Intensity Transfer Function; INCA – Input Catalogue Consortium; mas – milliarcsecond; NDAC – Northern Data Analysis Consortium; OTF – Optical Transfer Function; PFOV – Preceding Field of View; SM – Star Mapper; RGC – Reference Great Circle; RGO – Royal Greenwich Observatory; TDAC – Tycho Data Analysis Consortium. 1. Introduction With the rapid advances of astrophysics over the past 100 years there has been an increasing need for improvements in the quality and quantity of the most fundamental astronomical observations: the measurements of positions and trigonometric distances, or parallaxes, of stars. The practical limits for such observations from the Space Science Reviews 81: 201–409, 1997. c 1997 Kluwer Academic Publishers. Printed in Belgium. 202 F. VAN LEEUWEN ground seemed to have been reached, however, and these limits were no longer providing adequate constraints for the researches in stellar structure, stellar evolution and the Galactic distance scale, which form the basis for the extragalactic distance scale and cosmology (see also Kovalevsky, 1984). Moreover, ground-based aquisition of astrometric data is a time consuming and laborious process, relying on long-term investments. It is the kind of work for which there is, sadly, little space or provision in today’s world. More recently, various techniques have been developed that do allow very accurate relative astrometric measurements to be obtained from the ground, but still only in very small quantities (Zacharias, 1997, Harris et al., 1997). 1.1. PARALLAXES The measurement of parallaxes from the ground has always been very difficult. The traditional method made use of photographic plates, which, in order to get the maximum parallactic displacements, had to be taken at far from ideal conditions, away from the meridian and close to Sun-set or Sun-rise. In addition, the parallactic displacements could only be determined relative to a selection of reference stars on the same plate. The photographic plates used for this kind of work were exposed with very long focal length instruments, covering only a very small area of the sky on a plate. For example the refractors of Allegheny, McCormick, Sproul, Yale and Yerkes, which were used for many parallax determinations, all had focal lengths of 10 to 20 m. In a small area of the sky, all stars share the same parallactic displacement direction, while the sizes of the displacements depend on the distances of the individual stars. In the plate reductions, the reference stars were therefore also subject to the parallactic displacements, and this led to the fact that ground based parallax determinations always produced a measurement of the parallax relative to the average parallax of the background stars. An estimate of this background parallax gave an approximation for the correction to an absolute parallax. Such estimates were, however, always difficult and uncertain. It was not uncommon to see different parallax determinations for the same star with much larger differences than the precisions indicated for these determinations (for further details and references see Jenkins, 1952, Jenkins, 1963, Upgren, 1977, Gliese & Upgren, 1990, van Altena et al., 1995). Accuracies obtained for ground based parallaxes were at best of the order of 0.008 arcsec. More recently, using CCD technology, significantly higher reliability (to a level of 0.002 arcsec) has been obtained for small numbers of ground based parallax determinations for faint stars (Monet et al., 1992). The accuracies of the parallax determinations can be translated into accuracies of distance moduli (but note remarks made in Section 11) as well as in a distance range within which distance moduli can effectively be determined. Relative errors on the distances of less than 10 per cent were only reached for a very small number of objects, and even such accuracies gave uncertainties in the distance moduli for those stars at a THE HIPPARCOS MISSION 203 level of 0.1 to 0.2 mag, compared with accuracies of photometric measurements at a level of 0.001 mag. The accuracies of the parallaxes limited the distance range for useful measurements to a sphere with a radius of around 12 pc (0.2 mag accuracy in distance moduli, some 250 to 300 stars) to 30 pc (0.5 mag accuracy in distance modulus, some 4000 stars) from the Sun. The European Space Agency’s astrometric satellite Hipparcos has improved this situation dramatically. The parallax accuracies were improved by a factor 6 to 8 with respect to average ground-based accuracies, leading to a 100 to 1000 times larger volume within which significant parallaxes could be measured, and to a very significant increase in the numbers of stars available for distance determinations. Within its limiting magnitude, parallax measurements with accuracies better than 10 per cent were obtained by Hipparcos for around 20 000 stars. 1.2. ABSOLUTE POSITIONS Absolute positions of stars used to depend completely on meridian observations taken over long periods of time. This determined the stellar positions and hence a reference frame for a given epoch as well as the stars’ proper motions, all linked to the definitions of the Earth rotation (time) and precession of the equinoxes. The most recent such reference frame is known as J2000, and its realization is provided by a selection of 1535 bright stars for which the astrometric parameters were presented in the FK5 Catalogue (Fricke et al., 1989). Its positions and proper motions were derived from meridian observations that had been obtained with a wide range of manually operated instruments and many different observers. Some of the problems with the FK5 positions were exposed in the results of the first automated meridian instruments (Morrison et al., 1990), and the overall systematic errors in positions and proper motions became very clear after the comparison between FK5 and Hipparcos positions (ESA, 1997). In particular, in the southern hemisphere there were local discrepancies of several tenths of an arcsec. This is an unacceptable level for comparisons between radio maps and optical images obtained with the Hubble Space Telescope. Positional accuracies for radio maps can reach levels of 1 to 10 mas, and optical images from the Hubble space telescope have a resolution of around 0.1 arcsec. An error of a few 100 mas in positional alignment can obscure the proper physical explanation of the observations. 1.3. THE EMERGING OF HIPPARCOS In the 1960s P. Lacroute in France developed an alternative technology for measuring parallaxes, with particular emphasis on obtaining absolute rather than relative values. The basic idea was to combine the images from two areas on the sky onto one detector, and to do this at different times of the year. The parallactic displacements for the two fields would be uncorrelated and therefore allow the elimination of the parallax zero point problem. It was soon realized that such technology could 204 F. VAN LEEUWEN not be employed successfully from the ground, as it required a very well determined and stable angle between the two fields of view, and viewing directions from the ground are affected by a variable refraction. The first ideas of implementing this technique in a space mission were submitted in France in 1966 and 1967. The implementation appeared too complex at that time, in particular with respect to the construction of the beam combiner, and in 1970 further studies were stopped (see Lacroute, 1982). The basic idea, however, survived and with the input of several new ideas (many originating from E. Høg, Copenhagen Observatory), the concept of a space astrometry mission became more and more realistic. A meeting on space astrometry organized in 1974 in Frascati by the European Space Agency, ESA, marked the start of what was to develop into the Hipparcos project. Its official status as a phase A study by ESA was obtained in 1976, and it was approved as an ESA project in 1980. Hipparcos is both an acronym and a dedication to the Greek astronomer Hipparchus. As an acronym it stands for “HIgh Precision PARallax COllecting Satellite”. Hipparchus, who lived in the 2nd century BC, is generally associated with compiling the first catalogue of star positions from records of Babylonian observations as well as from some observations probably done under his auspices. At a very early stage of the first studies the aims of the mission were set. The astrometric data would be obtained with a scanning satellite over a nominal 2.5 year period, using an input catalogue of approximately 100 000 stars evenly distributed over the sky (Høg, 1979). The 2.5 year period was required for a reliable disentanglement of proper motions and parallaxes. The even distribution and the density of stars were needed to provide distortion free rigidity to the positional reference frame. The accuracy of the positions and parallaxes should be 0.002 arcsec, and for the proper motions 0.002 arcsec yr,1 , in order to make a sufficient impact with respect to the ground-based data. These clear aims were a driving force behind many of the design features of the spacecraft and payload. They required a mechanically (damping of vibrations) and thermally very stable system, and optics polished to very high precision. The original idea of Lacroute was incorporated in the two fields of view, projecting on the same detector. The critical element in the optical system was the “beam combiner”, a polished mirror which was cut in two halves and glued back at an angle, equivalent to half the angle between the areas on the sky as seen through the two fields of view. This larger angle, very close to 58 is referred to as the basic angle. The stability of the beam combiner determined the stability of the basic angle, an important criterion in the design of the payload. This was not the first time an optical element cut in halves was used for triangularization of the sky: in the early 19th century Bessel (1841) used for obtaining the most accurate relative positional measurements of those days a heliometer, an instrument in which the objective was cut in two. The intersection of the two halves of the objective was aligned with the great circle through a pair of stars. One THE HIPPARCOS MISSION 205 element of the cut lens could be shifted such that the images of the two stars would coincide. The measurement of the shift produced, like Hipparcos 150 years later, a one dimensional positional measurement, the angular separation of the stars. The measurements by Bessel of 61 Cyg provided the first reliable determination of the parallax of a star. His measurements on stars in the Pleiades cluster formed the first epoch of the 19th century astrometric studies of this cluster, which led to the distinction between field stars and cluster members (Lagrula, 1903) and the discovery of the colour magnitude relation for stars by Hertzsprung (1907). Hipparcos returned to the Pleiades cluster, and determined its distance (van Leeuwen & Hansen-Ruiz, 1997). A future astrometry mission, such as described in the GAIA concept (Lindegren & Perryman, 1995), will be able to determine the 3-dimensional structure of the cluster (van Leeuwen, 1995). The detector system for Hipparcos was in many ways similar to that which is used in meridian instruments. It is based on photoelectric measurements in conjunction with a slit system (Høg, 1960, Lindegren, 1979b). The purpose of both types of instrument is the same, i.e. the determination of an accurate transit time. In the case of the meridian telescope the images pass through a set of slits as a result of the Earth’s rotation; for the Hipparcos satellite, the passages were the result of the rotation of the satellite. The near synchronicity of measurements obtained in the two fields of view together with the scanning of the satellite made it possible to obtain a very accurate calibration of the basic angle, and thereby transforming very precise relative separations in the combined fields of view to very precise large angular separations on the sky. And this was the original aim of the proposals by Lacroute, which made it possible to determine absolute parallaxes. A requirement for the calibration of the basic angle was a smooth rotation of the satellite. In order to simplify the effects of solar radiation torques, the outer structure of the satellite was kept, in as far as possible, free from protruding structures. The satellite’s attitude control, which monitored and corrected the rotations of the satellite axes, was obtained through gyro and star mapper measurements for control and by cold-gas thrusters, producing very short bursts that changed rotation rates, for corrections. The measurements of large angles over the sky provided the means to build a very rigid positional frame. This frame was, however, “free floating”, not connected to other independent reference points. This is different from ground-based measurements which are always linked to the Earth’s rotation parameters: the rotation phase and the orientation in space of the rotation axis. The Hipparcos reference frame had to be “attached” to another very accurate earthbound reference frame in order to transform its precise positions into accurate positions. Several special programs were set up for this purpose. Most of these programs aimed at linking stellar positions to those of extragalactic sources. Very accurate positions of radio sources, including a small number of radio stars of which the optical counterparts were included in the Hipparcos catalogue, were obtained by means of large interferometer arrays (Lestrade et al., 1992, 1995, 206 F. VAN LEEUWEN Morrison et al., 1997a). The International Celestial Reference System (ICRS) was established on the basis of these radio positions and proper motions. A complete list of the radio sources used to define the ICRS was presented by Kovalevsky (1996). The ICRS is inertial through its linking to extragalactic radio sources and provides a link to the Hipparcos catalogue through a small overlap in the objects. Other methods, linking the Hipparcos proper motions and positions to the extragalactic reference frame using photographic surveys (Brosche & Geffert, 1988) or through the use of the Hubble Space Telescope (Hemenway et al., 1988) have also been used. The Hipparcos reference frame, after careful transformations, is now the optical realization of the ICRS (Lindegren & Kovalevsky, 1995, Kovalevsky et al., 1997, Section 8.7.4). 1.4. DATA HANDLING The handling of the data reductions was considered a major problem in the early developments. All observations were interconnected, and it appeared that the whole system needed to be solved in one go from the original phase estimates for the stellar transits to the astrometric parameters. This was far beyond the possibilities of the available computer technology. In 1976, L. Lindegren (Lund) designed a reduction system in which the solution was split into three parts, each of which was possible to handle. There was a theoretical small loss of accuracy associated with some approximations that had to be made, but these methods made Hipparcos possible in terms of data reductions. ESA, recognizing both the challenges in the hardware design and the difficulties in the data reductions, decided when the project was approved in March 1980, that all data reductions should be handled by the astronomical community in Europe. It was also decided, that, as there was no independent material available with which to compare Hipparcos results, it would be beneficial to the project if more than one group carried out the full data reductions. Two consortia applied for and were entrusted with the data-reductions responsibility: the FAST consortium, based in France, Italy, Germany and the Netherlands, and led by J. Kovalevsky, and the NDAC consortium, based in Sweden, Denmark and the UK, led initially by E. Høg and later by L. Lindegren. A third consortium, led by C. Turon, took the responsibility for the preparation of the Input Catalogue. 1.5. PRE-LAUNCH DEVELOPMENTS After the approval of the project, a second experiment was proposed by Høg et al. (1982). Using the signal obtained from the star mapper detectors, whose primary purpose was to help keep control of the satellite attitude, a complete survey of the sky down to magnitude 11 could be obtained. By splitting the star mapper light into two channels, roughly resembling the V and B bands from the Johnson system, the survey would be supplemented with photometric information that would also be THE HIPPARCOS MISSION 207 useful to the main mission. This parallel mission was named Tycho, after the Danish 16th century astronomer Tycho Brahe (1546-1601), who not only discovered the nova in Cassiopeia in 1572, but also compiled the most accurate pre-telescope era catalogue of positions for some 800 stars, using a very large quadrant. A fourth consortium, TDAC, based in Denmark and Germany, took the responsibility for the Tycho data reductions and was led by E. Høg. The Tycho astrometry was of lower accuracy than that obtained with Hipparcos, but the completeness and the density of stars in the survey (an average of 25 stars per square degree) offered many possibilities for further research. The years up to 1989 were spent by the data-reduction consortia on the development of the computer codes for the data reductions and the interfaces with the data delivery. Extensive technological developments took place at the same time, but these are not covered by the current review (see instead Volume I of ESA, 1989 and Volume 2 of ESA, 1997). Several experiments were done using simulated satellite data created at ESTEC, RGO and CERGA. The data-reduction results of these simulated data provided the input for the first comparison exercises, and the first quality checks on the data reductions. These comparisons were going to be a regular feature of the data reductions during the mission, and contributed much to the outstanding quality of the final data. 1.6. POST-LAUNCH DEVELOPMENTS On August 8, 1989 the satellite was launched successfully into a geostationary transfer orbit by an Ariane 4 rocket (launch V33). Soon after, problems started to develop when the apogee boost motor (ABM) failed to react on firing commands and left the satellite stuck in this highly elliptical orbit. This caused many problems: three ground stations were required to keep sufficient contact with the satellite (see Figure 1); during the low perigee passages no contact was possible and no observations could be made, often leading to loss of attitude control by the satellite; the increased torques working on the satellite close to the perigee passage led to short periods of relatively high usage of the cold-gas thrusters, shortening the lifetime of the satellite; in this orbit, the satellite crossed the Van Allen radiation belts about once every five hours, leading to a gradual deterioration of solar panels and electronic equipment. Despite pessimistic estimates in the autumn of 1989 (see for example Waldrop, 1989, Neto, 1989) about the possible survival of the satellite in this environment, it operated for 3.5 years and obtained results about 1.5 to 2 times better than the original aims (Perryman et al., 1997a). Radiation damage to the on-board computer and the gyro electronics resulted in the termination of the observations in March 1993 and of the mission in August 1993. Behind the ultimate success of the mission was the enormous dedication from the ground staff at ESOC (led by D. Heger) and the data-reduction consortia, the support of NASA who made a receiver available at Goldstone, and last but not least the commitment of the Hipparcos Science Team led by the Hipparcos project scientist 208 F. VAN LEEUWEN Figure 1. The position of the satellite in the final orbit plotted relative to a stationary Earth. Points in the orbit are drawn at equal time intervals. The just over 9 successive orbits shown are numbered. Orbit 10 almost coincides with orbit 1. The ground stations are indicated by their initials: Odenwald, Perth, Kourou and Goldstone. For each ground station the horizon is drawn as a light grey line, showing the limited visibility of the satellite under various combinations of ground support. M.A.C. Perryman (see Table I). The close collaboration between the data reduction consortia and the ground staff at ESOC led to rapid recognition of any developing problems (first look analysis, Schrijver, 1985) and to improved calibrations for payload and spacecraft (gyro and torque calibrations done at RGO). The RGO was also closely involved in developments to keep the satellite in operation despite a break-down of all gyros. As a result of the additional ground stations and the ever improving ability at ESOC to recover the satellite’s attitude, the final data return was still approximately 65 to 70 per cent. 209 THE HIPPARCOS MISSION Table I The Hipparcos Science Team during the mission operations phase. U. Bastian P.L. Bernacca M. Crézé F. Donati M. Grenon M. Grewing E. Høg J. Kovalevsky F. van Leeuwen L. Lindegren H. van der Marel F. Mignard C.A. Murray M.A.C. Perryman R.S. Le Poole H. Schrijver C. Turon Heidelberg Asiago Strasbourg Torino Genève Tübingen Copenhagen Grasse Cambridge Lund Delft Grasse Eastbourne ESTEC Leiden Utrecht Paris-Meudon (D) (I) (F) (I) (CH) (D) (DK) (F) (UK) (S) (NL) (F) (UK) (NL) (NL) (NL) (F) from 1993 from 1989 from 1983 from 1981 from 1982 from 1983 from 1981 from 1981 from 1986 from 1981 from 1987 from 1991 from 1981 from 1981 from 1981 from 1990 from 1981 Table II Orbit characteristics. Element Mean value Change/yr Variation Units 38340 24582 0.7196 6890 0.37 -20.4 -8.2 0.000 -3.0 4.5 2.5 0.005 140 0.07 s km km /day Period Semi-major axis Eccentricity Perigee Precession 1.7. THE OPERATIONAL ORBIT Using the thrusters and fuel that was intended for moving the satellite to its correct geostationary position, the perigee height was increased to move the satellite away from the outer layers of the Earth’s atmosphere. The remaining fuel was drained to avoid sloshing. An orbital period was obtained such that regular patterns would occur in ground-station usage and occultations. The final orbital period was close to 5 rotational periods of the satellite, and 9 orbital periods spanned almost exactly 4 days (see Figure 1). The average orbital parameters are summarized in Table II. The data reduction requirements called for instantaneous knowledge of the satellite position to 1.5 km and its velocity to 0.2 m s,1 (Perryman et al., 1992). 210 F. VAN LEEUWEN Much more than would ever have been the case for the geostationary position, the conditions of the satellite orbit determined the handling of the satellite and the data reductions. With a guaranteed break in the observations of some 2 hours around perigee, the “data set” units were effectively defined, giving the maximum length of time for a wide range of calibration and data reduction processes. Given the orbital period of just over 10 hours, the average data set covered some 6 to 8 hours of observations. This was a shorter period than originally foreseen for the great-circle reduction process, which was therefore slightly more vulnerable to instabilities caused by the closeness of the basic angle of 58 to a 6th harmonic at 60 . Such instabilities were controlled in the penultimate stage of the data reductions, the sphere solution (see Section 8.6). The orbits were numbered sequentially, starting at the first scientific data obtained on November 5, 1989. The total mission covered a time-span of 2768 orbits, equivalent to 1230 days. Observations were obtained during 2341 orbits. In a few cases successfully reduced data was obtained by only one of the reduction consortia. This applied for 94 orbits, most of which concern short stretches of data. An 11-week gap in the data occurred from orbit 2274 till orbit 2446, when the satellite was being prepared for operations using two instead of three gyros (the period indicted by ‘e’ in Table V and Figure 18). Another consequence of the operational orbit were the much extended eclipse periods and eclipse lengths. With the satellite moving to within 400 km from the Earth’s surface, eclipses could last for almost 2 hours (Figure 2), compared to a maximum length for the nominal mission of 72 min. This put severe demands on the power supply, with the time between eclipses only marginally long enough to reload the batteries. The satellite also cooled down very significantly during such periods (see Section 3.1), which was observed in various instrument calibration parameters despite the very strict thermal control of the payload. The eclipses also caused problems with the reconstruction of the satellite attitude due to the changes in the solar radiation torques working on the satellite (Figure 37). Two types of torques, the gravity gradient and the interaction between the Earth’s magnetic field and the magnetic moment of the satellite, increase by ,3 , where is the distance between the satellite and the centre of the Earth. These torques would have been insignificant in the geostationary orbit, but now played an important role (see Section 8.2.2). Occultation periods, when the image of the Earth or the Moon came too close to one of the fields of view, were also on average longer than would have been the case in the geostationary orbit. This led in some cases to extended interruptions around perigee, when the perigee passage was preceeded or followed by a long occultation. With the two fields of view, occultations usually came in pairs, separated by about 20 minutes. Further details on how the occultations affected the main-mission data reductions can be found in Sections 8.2 and 8.5. D D THE HIPPARCOS MISSION 211 Figure 2. Reflections of temperature changes in the spacecraft. Top graph: lengths of eclipses; Middle graph: exposure factor (see text); Bottom graph: drift rate of the on-board time (OBT) measured relative to the ground-station time (GST). 212 F. VAN LEEUWEN 1.8. DATA RELEASE The results of the Hipparcos mission were made public in June 1997 in the form of printed catalogues (for selected results only) and on ASCII CD-ROM for all relevant data (ESA, 1997). Included in the printed catalogue is the documentation of the data reduction processes and selected data files. A set of four letters to Astronomy & Astrophysics described briefly the main-mission astrometric results (Perryman et al., 1997a), the Tycho astrometric and photometric results (Høg et al., 1997), the photometric data (van Leeuwen et al., 1997a) and the double star investigations (Lindegren et al., 1997). Before this release date, selections of the data had been made available on request to those involved in the preparation of the Input Catalogue and the data reductions (September 1996) and to those who made observing proposals in 1982 (release date January 1997). A symposium in Venice in May 1997 rounded off the project and presented the first scientific results (Perryman & Bernacca, 1997). In 1994, developments started on a possible follow-up mission, building on the expertize obtained from Hipparcos and on much improved technological possibilities. The ROEMER mission proposed to utilize CCDs for drift scan in a Hipparcos-like satellite (Høg, 1993), (Høg & Lindegren, 1994), thus improving the light collecting efficiency of Hipparcos by a factor 105 . This was further developed in the GAIA proposal, using interferometric techniques, which should be capable of improving the Hipparcos results by about a factor 100 in accuracy, and 500 or more in numbers of stars (Lindegren & Perryman, 1995, Perryman & van Leeuwen, 1995). 2. Possibilities of ground-based astrometry In this section the Hipparcos aims and results are compared with the achievements and possibilities of ground-based astrometry. Although some aspects of ground-based astrometry have become less meaningful with the availability of the Hipparcos and Tycho data, in most cases, the Hipparcos and Tycho data will lead to a significant enhancement of the possible results of ground-based astrometry through providing a well defined high accuracy positional reference frame. At the same time, the use of ground-based data can provide an extension to the Hipparcos and Tycho catalogues to much fainter magnitudes (the determination of accurate positions for large numbers of very faint objects in the International Celestial Reference System, ICRS) and an increase in the base line used for the proper motion determinations (providing a better distinction for orbital disturbances to Hipparcos proper motion and parallax determinations). 2.1. MERIDIAN ASTROMETRY Traditional meridian astrometry, where the goal was to maintain the positional reference system, is the only technique that has become largely redundant with THE HIPPARCOS MISSION 213 the release of the Hipparcos data. The definition of the celestial reference system using automatic meridian telescopes was accurate to a level of 0.04 to 0.1 arcsec, compared to the Hipparcos reference frame (at epoch J1991.25) to a level of at least 0.001 arcsec. Due to its short time-basis, this accuracy will deteriorate relatively quickly, but it will take some 50 to 100 years before the Hipparcos accuracy has become comparable with the meridian-circle data. It is hoped that before that time comes, at least one follow-up astrometric space mission will have substantially extended this limit. In the case of the proposed GAIA mission it would take many thousands of years for the absolute positions to deteriorate to the level of 0.1 arcsec. The future for the meridian instruments, and in particular the automated ones, is now in extending the positional reference frame from the Hipparcos and Tycho limiting magnitude of 11, to magnitude 15 or 16, and thus provide a link between the Hipparcos and Tycho reference frames and the positional measurements of magnitude 13 to 21 objects on photographic plates or CCD frames. This can be done by putting a CCD detector in the focal plane and taking images or making drift scans (Stone, 1997). 2.2. WIDE FIELD PHOTOGRAPHIC ASTROMETRY The main use of wide-field astrometry is as an intermediary in providing reference positions for faint objects, and as an instrument for large-scale proper motion studies (see for example Murray et al., 1989, Murray, 1989, Evans, 1992, Evans & Irwin, 1995). The possibilities for providing reliable reference positions will be very much improved with the release of primarily the Tycho data, as long as plates are available sufficiently close in epoch to the mean Hipparcos and Tycho epoch, J1991.25 (Robichon et al., 1995). This will then allow for a proper description of the plate transformation and reliable positions for fainter objects. Even with plates taken at larger epoch differences it will be at least partly possible to disentangle the proper motion effects from the plate distortions. A major difficulty is, however, caused by the magnitude range of Hipparcos and Tycho Catalogues in comparison with the magnitude range of a typical Schmidt survey plate. All Tycho images will normally be saturated on such a plate. This is where the newly equipped meridian circles can step in: they can provide the link from the magnitude range of Hipparcos and Tycho to the useful magnitude range of a survey plate. Very significant contributions in this field are also expected from electronic digital sky surveys, such as Sloan (Kron, 1995) and DENIS (deep near-infrared survey, Le Betre et al., 1995). In the use of survey plates for proper motion studies, the main contribution from the Hipparcos data comes in the form of defining the reference system for those proper motions, i.e. making them absolute rather than relative. Also, a better understanding of the distortions to which those plates are subjected can improve the comparisons between plates of different epochs and thus decrease the extent to which these distortions enter the proper motion determinations. These improve- 214 F. VAN LEEUWEN ments are all important for the study of Galactic dynamics. Thus, it is expected that Hipparcos and Tycho astrometry enhances the possibilities of ground-based wide-field astronomy, and can help extending the Hipparcos reference frame to fainter magnitudes. 2.3. SMALL-FIELD PHOTOGRAPHIC ASTROMETRY Small-field astrometry explores the large numbers of photographic plates taken with photographic refractors from as early on as the 1880s. Small-field photographic astronomy started in the 1850s with wet-emulsion photographs of the moon, and received a major boost with the work done at Paris Observatory by the brothers Paul and Prosper Henry in the 1880s. Their success in producing photographic doublets of flint and crown glass, demonstrated by many very good photographs of e.g. the Moon and the Pleiades cluster, led, at the end of the 19th century, to the Astrographic Catalogue initiative, which was later followed by the Carte du Ciel project. Both projects used refractors with a fixed focal length of 3.4 m and an objective diameter of 30 cm. It is interesting to notice that this first photographic sky survey is now being re-examined for the use as a first epoch to the Tycho data (Kuzmin et al., 1997). Later developments led to refractors with longer focal lengths and larger objectives. The great astrometric value of the plate material accumulated with these telescopes stems from the useful magnitude range (2 to 16), combined with a reasonable plate scale, field size and a large spread in epochs. All this is available thanks to the long term views of astrometrists during the first half of the 20th century. This kind of material has provided the most precise relative proper motions for stars in star open clusters, and more recently also in globular clusters, with precisions as high as 0.1 to 0.2 mas yr,1 , ten times better than Hipparcos and reaching to much fainter magnitudes (see for example Vasilevskis et al., 1979, and references in van Leeuwen, 1985). The combination of this material with the Hipparcos data is in principle possible, but hampered by distortions that occur in proper-motion reference frames derived this way (Tian et al., 1996). In addition, there are problems with comparing short time-base proper motions such as determined by Hipparcos with long time-base proper motions determined using ground-based data (Wielen et al., 1997), problems which will become more severe as improved positional accuracies are obtained over short epoch intervals only. In some studies small-field astrometry will for a long time to come be the main or only provider of high-precision astrometric data, most noticeably for measuring internal motions in globular clusters, where star densities are generally too high to handle for a scanning instrument. Small-field astrometry for the use of parallax determinations can also be improved through the much better possibilities for defining the parallax zero point. Thus, ground-based parallaxes can be brought closer to absolute measurements. THE HIPPARCOS MISSION 215 2.4. CCD ASTROMETRY This is a recent development in astrometry and early results have shown very good possibilities for determining parallaxes for fainter stars (Monet et al., 1992, see also Connes, 1979 and Lindegren, 1980 on aspects of potential accuracies of ground-based measurements). CCD frames are also being used to determine accurate positions of faint objects, connected through many links of CCD frames and photographic plates to the Hipparcos and Tycho catalogue positions. Examples of such applications were presented by Reynolds et al. (1995) and Caraveo et al. (1997). 3. The Hipparcos satellite The satellite was designed and built by an industrial consortium led by Matra Marconi Space in France and Alenio Spazio in Italy under contract to ESA. The satellite housed on its lower platform the spacecraft, which consisted of the hardware for running and controlling the instrument, and on the upper platform the payload, which contained the telescope and the detector chains (see Figure 3). The satellite was controlled from the ground-segment at ESOC in Darmstadt through the ground stations in Odenwald (Germany), Perth (Australia), Goldstone (USA) and Kourou (French Guyana). 3.1. THE SPACECRAFT The task of the spacecraft was to ensure the proper operation of the payload. This ranged from getting it in the right orbit (in which it was not very successful) through power supplies and handling all communications with the ground stations, to controlling the scanning of the satellite through thruster firings. A major part of the spacecraft platform was occupied by the apogee boost motor (MAGE 2 ABM) tank (see Figure 3), which was still full with solid fuel. Due to this, the total mass of the satellite during operations was considerably higher than was planned: 1100 kg instead of 640 kg. As a consequence, the satellite was less affected by solar radiation torques (the same torques resulted in smaller accelerations around the rotation axes), and the centre of mass for the satellite was shifted away from its nominal position by 72 cm. The thrusters for the z axis of the satellite were therefore no longer in a plane through the centre of mass of the satellite, and firings of the z axis thrusters produced consequently residual torques on the x and y axes. The latter effect allowed, using the reconstructed-attitude data, a determination of the position of the centre of mass for the satellite to an accuracy of about 1 mm. The spacecraft handled the control of the scan by means of 6 cold-gas thrusters, supplied from one of two cold-gas tanks. There were two thrusters each for the x, y and z axes, producing positive and negative changes to the rotation rates respectively. Thruster firings took place under control of the real time attitude determination 216 F. VAN LEEUWEN Figure 3. Exploded view of the satellite, illustrating the position of the payload and below the payload the spacecraft platform. program running on the on-board computer. They lasted in general for less than half a second (one quarter of an observational frame). The intervals between thruster firings varied from a few minutes to more than half an hour. Near perigee passages the thruster firings became more frequent and longer, compensating for the much increased gravity gradient and magnetic field torques. Thruster firings affected the data reductions by producing discrete changes in the satellite attitude, which THE HIPPARCOS MISSION 217 had to be properly provided for in the attitude reconstruction and the great-circle reduction. The power supply on-board the satellite came from 3 solar panels, backed up by batteries that were activated during solar eclipses. The total power consumption during normal operations was around 270 W, with at the start of the mission a power supply of 445 W by the solar panels, dropping to 360 W towards the end of the mission as a result of radiation damage to the solar cells. During Sun-pointing mode the power supply from the solar panels dropped due to the higher temperature, and towards the end of the mission the satellite could no longer fully function in Sun-pointing mode. The power subsystem provided a 50 V dc bus to the satellite power users. A detailed report on the solar array performance and degradation as a result of radiation can be found in Chapter 11 of Volume 2 of ESA (1997). Some equipment on the spacecraft platform was thermally controlled, but overall the conditions on the platform reflected the exposure of the satellite to sunlight. This was most clearly demonstrated by measurements of drift rates in the on-board time (OBT) that resulted from temperature changes. Times as provided by the OBT were referred to the much more accurate time-tagging of the data at the ground stations, and the evolution of the differences between the OBT and ground-station times (GST, corrected for satellite to ground-station and receiver delays) provided a measure of the OBT drift. The OBT was regulated by a quartz oscillator, situated on the spacecraft deck close to the outside cover of the satellite, roughly halfway between the apertures of the two fields of view. The long-term changes in the OBT drift were directly correlated to the exposure factor, defined here as the fraction of time of an orbit during which the satellite was exposed to the Sun, divided by the square of the distance d between the Earth and the Sun: x E = Torb , Tecl Torb 1 2 d ; (1) where Torb is the orbital period and Tecl the length of time of an orbit during which the Sun was eclipsed by the Earth. Both the ellipticity of the Earth’s orbit and the lengths of eclipses can easily be recognized in the drift of the on-board clock (Figure 2). There were smaller effects relating to the height of perigee (warming up of the spacecraft due to friction with the outer layers of the Earth’s atmosphere, see also Section 8.2.6) and the rotation of the satellite (for more details see ESA, 1997, Volume 3, Chapter 8). The communications from the spacecraft to the ground station were arranged through a continuous telemetry stream of 23.04 kbits per second, containing all science data, attitude-control-related data and housekeeping information such as temperatures. Telecommands from the ground station to the spacecraft were used to provide the instrument with a program star file, a listing of stars with predicted positions and magnitudes, that were expected to become visible in either of the two fields of view. This data was stored in a buffer on-board the satellite. The buffer could store the star data for at least 5 minutes of observations. 218 F. VAN LEEUWEN Table III Payload characteristics. Optics Telescope configuration Field of view Separation between fields Diameter of primary mirror Focal length Scale at focal surface Mirror surface accuracy All-reflective Schmidt 0.9 0.9 58 290 mm 1400 mm 6.8m per arcsec =60 rms (at 550 nm) Main detector Modulating grid Slit period Detector Photocathode Scale at photocathode Sensitive field of view Spectral range Sampling frequency 2688 slits 1.2074 arcsec (8.2m) Image dissector tube S20 3.0m per arcsec 38 arcsec diameter 375-750 nm 1200 Hz SM detector Modulating grid 4 slits perpendicular to scan 4 slits at 45 inclination Photomultiplier tubes Bi-alkali eff 430 nm, 90 nm eff 530 nm, 100 nm 600 Hz Detectors Photocathode Spectral range (BT ) Spectral range (VT ) Sampling frequency = = = = = 3.2. THE PAYLOAD The Hipparcos payload was built around an all-reflective Schmidt telescope, working in the visual part of the spectrum. It consisted of four main assemblies. two external baffles to protect the telescope from stray light; a carbon-fiber reinforced box containing the telescope assembly, consisting of three mirrors, and mounted onto the spacecraft through three titanium bipods; the focal plane assembly, containing all elements from the grid to the image dissector tube and star mapper photomultipliers. This assembly could be shifted to obtain an optimal focus position (see Figure 7); the service electronics units. The main characteristics of the payload are shown in Table III. The optical system consisted of a beam combiner, which also took the role of Schmidt corrector (providing the off-axis correction for spherical aberration), a flat folding mirror, and a spherical primary mirror (see Figure 4). The mirrors were made from lightweighted Zerodur blank, polished to =60 rms at a reference wavelength of = THE HIPPARCOS MISSION 219 Figure 4. Configuration of the Schmidt telescope. Light entered from the two baffle directions, and was brought together at the beam combiner, which was configured as a Schmidt corrector. The combined light was reflected from the flat folding and spherical primary mirrors onto the focal surface where the modulating grids were located. Figure 5. The grid assembly characteristics. 550 nm. The basic angle between the two fields of view had been chosen on the basis of rigidity tests for the great-circle reductions and the provision of sufficiently different parallax factors in the two fields of view. In the focal plane of the telescope was situated the grid assembly. The characteristics of the detector grid are shown in Figure 5. The optical block on which the 220 F. VAN LEEUWEN grid was written consisted of a piece of Suprasil 1, a high quality silica. The convex surface of the optical block matched the curved focal surface of the telescope. Its rear side for the star mapper grids was flat, and for the main grid was ground with a radius of curvature of 213.8 mm, which made it serve as a field lens. The surface of the optical block had to be of extreme smoothness, as it was to support the detector grids with sub-micron accuracy. After polishing and cleaning, a layer of chromium was sputtered on the surface, followed by the deposition of a layer of electron-sensitive resist. The main grid and the star mapper grids were engraved in discrete “scan fields”, with 46 168 of these scan fields covering the main grid field of view. The method of engraving these grids meant that when projected orthogonally on a plane normal to the optical axis of the optical block, these grids are strictly rectangular, but while projected on the curved surface of the optical block, they are slightly distorted (see Section 8.4). The spacing of the grid lines was 8.20 m, and their width 3.13 m. The grid was made to very high specifications, with the rms variation in the main grid parameters at a level of 0.010 m or less. A full specification of the main grid and the star mapper grid is presented in Tables 2.5 and 2.6 of Volume 2 of (ESA, 1997). Figure 6 demonstrates the very high precision of the main grid as measured on the ground. The optimal focus of the telescope was determined by maximizing the modulation amplitude resulting from the modulating grid. It turned out, however, that the optimal focus for the two fields of view were different, and a compromise value halfway between the two best focus positions was adopted (see Figure 7). As a result of the different focal positions, the effect of the telescope focus drifting during the mission was noticeably different for the two fields of view, with the preceding field of view (PFOV) moving towards a better focus, while the following field of view (FFOV) was moving away from it (see for example Figure 20 or Figure 47). The payload units were kept under very strict thermal control in order to minimize changes in the optical system. The thermal stability of the payload can be shown from the evolution of the basic angle as well as for various other calibration parameters. Only periods of very long eclipses and excursions into Sun-pointing mode can be recognized, but none of the other features shown in Figure 2 are reflected in Figure 8. The range of short-term variations was of the order of 1 mas, for long-term variations the range was about 20 mas. Only as a result of Sunpointing mode conditions and failures in the thermal control electronics did larger variations occur. 4. The Input Catalogue For a complete description of the Hipparcos Input Catalogue the reader is referred to Volume II of ESA (1989) and Turon et al. (1992a). A summary of the properties THE HIPPARCOS MISSION 221 Figure 6. The medium-scale irregularities of the main grid in the scan direction as determined through ground-based measurements. The geometry of the grid corresponds to the configuration shown in Figure 5. The scale in the upper figure is such that one horizontal scan-field unit corresponds to 40 nm. The lower figure shows the mean offset per slit line in the scan direction (solid line) and the standard error of the offsets (dotted line). of the Input Catalogue is provided by Turon et al. (1995). Here only some essential points are reviewed. The aim of the Hipparcos Input Catalogue was to provide the Hipparcos mission with a fixed selection of approximately 100 000 stars, evenly distributed over the sky (in as far as possible), for which positions at the epoch of observation are known to an a priori accuracy of better than 1.5 arcsec, and 222 F. VAN LEEUWEN Figure 7. Determination by ESOC of the optimal focus in the two fields of view for stars of different colours, by maximizing the modulation amplitude. The optimal focus position for the two fields of view differed, and a halfway position was adopted as the optimal focus. Figure 8. The evolution of the basic angle as determined by NDAC. The feature near day 420 is associated with the long eclipses around that time, and an overall cooling down of the satellite. with sufficiently well known photometric properties. At the same time, the Input Catalogue was supposed to incorporate stars requested in observing proposals made in 1982. The preparation of the Input Catalogue was the responsibility of the INCA consortium, led by C. Turon. 223 THE HIPPARCOS MISSION Table IV The Scientific Proposals Selection Committee, 1982-1988. A. Blaauw (chair) J. Dommanget W. Gliesey M. Hack E.P.J. van den Heuvel C. Jaschek J. Lequeux P.O. Lindblad A. Maeder P.E. Nissen B.E.J. Pagel A. Renzini C. de Vegt P.A. Wayman R.Wielen 4.1. OBSERVING PROPOSALS Following an invitation by ESA at the end of 1981, some 200 proposals for objects to be observed were submitted. They covered 214 000 different objects, and it was clear that a fairly large number of these would have to be rejected. The proposals covered a wide range of topics, from luminosity calibrations of specific types of stars through distance calibrations for star clusters and associations to Galactic dynamics. A selection from these proposals was made by the Scientific Proposals Selection Committee (see Table IV). A first selection was made by ranking those proposals which were possible to be observed, and priorities were assigned to individual objects. It was realized that faint objects required much more observing time than bright objects, and that an increase in the number of faint objects included in the selection would lead to an overall decrease in the total number of objects that could be included in the catalogue. The distribution of stars over the sky resulting from this first selection indicated that a large amount of work still had to be done to achieve a sky and magnitude distribution better suited to the needs of the Hipparcos mission. 4.2. SELECTION MECHANISMS As a first step, proposers were informed of, and asked to comment on, the recommendations by the selection committee, and to examine the consequences of these recommendations. Based on the responses of the proposers, the selection committee reviewed their recommendations. Also the overall contents of the Input Catalogue were reviewed and the concept of a magnitude-limited complete survey was introduced. Studies by the data reduction consortia provided details on the feasibility of double star observations and disturbances by nearby bright stars. Meanwhile, groups within INCA were improving the mission simulations, which allowed for every selected object to obtain an estimate for the expected accuracy of the astrometric parameters. Different algorithms, giving the observing time as a function of stellar magnitude, were tested, allowing decisions to be made about the total number of stars that could be retained in the catalogue. Other groups in INCA were optimizing the observing lists for minor planets and variable stars, and unifying proposals in dense regions of the sky. Final selections were often made 224 F. VAN LEEUWEN based on the results of the simulation software. The final results were presented to the selection committee in the form of detailed statistics and performances for each proposal, allowing the selection committee to verify that their recommendations had been implemented satisfactorily. 4.3. PREPARATION OF THE CONTENTS OF THE INPUT CATALOGUE For all stars appearing in the preliminary selections accurate coordinates and photometric properties had to be obtained. In many cases this meant new observations, using plate measurements, meridian telescopes and photometers (Jahreiß et al., 1992). In particular for many variable and large proper motion stars the coordinates provided by the proposers were often of very poor quality, and it is also mainly among these stars that the final Input Catalogue contains a small number of misidentifications. The SIMBAD data base at the CDS in Strasbourg was extensively used in these preparations. Photometric, spectroscopic and radial-velocity data were collected from the literature for the selected objects. When no photometric data was available, and the star was too faint to expect Tycho observations of sufficient accuracy, stars were incorporated in ground-based observational programs of the Danish, Dutch and Swiss telescopes on La Silla, ESO, providing Strömgren, Walraven and Geneva photometric measurements respectively (Grenon, 1988, Grenon et al., 1992). A collaboration with the AAVSO was set up to assist with the observations of Miras and other large-amplitude variable stars (Mennessier & Baglin, 1988, Mattei, 1988). Special groups were set up for the aquisition of radial velocities for stars in the Input Catalogue (for further details see Burnage et al., 1988, Udry et al., 1997, Fehrenbach et al., 1997). 4.4. PROPERTIES OF THE INPUT CATALOGUE The Hipparcos Input Catalogue was published in printed form and on CD-ROM (ESA, 1992). As was described in the preceding sections, the final Input Catalogue was a compromise between observing proposals, mission requirements and limiting magnitude. The contents are summarized in Figures 9 to 11. 4.4.1. Completeness The Input Catalogue is nominally complete down to the following limits: V V 7 9 + 1 1 sin j j for spectral types earlier or equal to G5 7 3 + 1 1 sin j j for spectral types later than G5, : : b : : b where b is the galactic latitude. Within these limits, 55 000 objects were selected from the SIMBAD data base. Among the very few bright objects missing are components of very close systems of bright stars such as in the Orion Trapezium cluster. For fainter magnitudes the selected objects had to fulfil one of two requirements: THE HIPPARCOS MISSION 225 Figure 9. The cumulative numbers of stars in the Input Catalogue as a function of distance based on the parallaxes in the final catalogue. The grey line represents an increase by r3 , showing the differences between the numbers of stars included and the total number of stars expected to be present within a given distance from the Sun (not corrected for density gradients). Figure 10. The distribution of magnitudes in the Input Catalogue as based on the mission determinations of magnitudes. The grey line represents the completeness limit, showing the Input Catalogue as essentially complete down to mag 7. being a selected object from one of the approved observing proposals or assisting in creating a homogeneous coverage of the sky. Thus, a substantial number of objects in the Input Catalogue was finally selected mainly for reasons of the rigidity of the final astrometric reference frame. 4.4.2. Double stars The double stars included in the Input Catalogue were to some extent selected on the basis of separation and magnitude difference, considering the possibilities 226 F. VAN LEEUWEN Figure 11. The distribution over the colour index (V , I )C . as well as the limitations of the Hipparcos main detector (Dommanget, 1985, Dommanget, 1988). In connection with the preparation of the Input Catalogue, a new double star reference catalogue was created with specific emphasis on astrometric properties: absolute positions and proper motions (Dommanget & Nys, 1994). Some 12 000 double stars were included in the Input Catalogue, and of these 9500 were possible to resolve on the main grid, having separations of at least 0.1 arcsec and magnitude differences of 4 mag or less. The target positions for these objects were the individual components (for separations beyond 10 arcsec), or else the photocentre or the brightest component of the system. 4.4.3. Variable stars Many pulsating variable stars, from Sct to Mira types, regular, semi regular and irregular, were included in the Hipparcos Catalogue for the purpose of luminosity calibrations, either directly through the measurement of parallaxes, or indirectly through the use of proper motions (Mennessier & Baglin, 1988). A small number of discoveries of pulsating variables were made among other stars in the catalogue, using the Hipparcos main mission or the Tycho epoch photometry (Turon & van Leeuwen, 1995, Bastian et al., 1996). Several hundred eclipsing binaries were also included, and here the Hipparcos discoveries almost doubled the number of cases in the Input Catalogue (see for example van Leeuwen & van Genderen, 1997). 4.4.4. Identification problems The main source of identification problems was the very inaccurate coordinates provided in some observing proposals or available in the literature. This applied primarily to two types of objects: faint variable stars and large-proper-motion stars. Among those stars misidentified in the Input Catalogue and therefore not observed THE HIPPARCOS MISSION 227 Figure 12. The heliotropic angles ( , , , and ) in which the satellite’s nominal attitude was described, and their relation to the satellite axes (x, y, z). The plane of the ecliptic is indicated by ‘E’, the scanning circle by ‘G’. The two fields of view (PFOV, FFOV) and the basic angle are also shown. by the main mission are the variable stars GQ Ori (a Cepheid) and SS Cyg. One of the consequences of misidentification was the use of the wrong colour index in the data reductions, which affected primarily the main-mission photometric results (see Section 9.3). 5. The scanning law The scanning law (see Figure 12) was designed to describe a homogeneous coverage of the sky. Restrictions on the scanning law came primarily from the brightness of the Sun and from the requirement to provide approximate predictions of the solar radiation torques, acting on the satellite, for the on-board attitude control. This required a near constant angle between the direction to the Sun as seen from the satellite, and the orientation of the satellite’s z axis. The solution that was adopted described the scanning law in ecliptic coordinates, with the z axis of the satellite precessing under an angle of = 43 around the direction to the Sun ( , see Figure 12). The precession rate for the z axis was 6.4 revolutions per year (_ 6:4_ ). The satellite was spinning around the z axis at a rate of 11.25 revolutions per day ( _ = 168:75 arcsec s,1). The precession rate was adapted to the ellipticity of the Earth’s orbit, to provide a smooth and homogeneous scan. Almost the entire sky was scanned in at least two directions every half year. Figure 13 228 F. VAN LEEUWEN Figure 13. The scanning motion of the Hipparcos satellite on the celestial sphere, shown in ecliptic coordinates. The top figure shows the path of the spin axis between 22 May 1990 and 23 September 1991. The scan direction is indicated by the arrows. The bottom figure shows one reference great circle for each orbit (5 revolutions of the satellite) for the period between 22 May and 14 July 1990. The actual scanning was 5 times denser. shows the sky coverage for a 2-month period and the path of the z axis for a 16month period in the beginning of the mission. Figure 129 shows the scan-coverage over a 6-month period. On a few occasions, the satellite returned to the Sun-pointing mode (with the z axis pointing towards the Sun), which was the fall back position for a serious failure in the satellite attitude control. The thermal equilibrium status of the satellite was disturbed at every such occasion, causing problems with some of the instrument parameter calibrations (instrument parameters were often found to evolve rapidly shortly after moving to Sun-pointing mode, while being assumed constant in the calibration models). In the main-mission epoch photometry all data acquired during THE HIPPARCOS MISSION 229 Sun-pointing mode observations is flagged as possibly suspicious. Main-mission astrometric data obtained in Sun-pointing mode did not contain any information on the stellar parallaxes, as in this configuration all parallax displacements were perpendicular to the scan direction. The peculiarities of the scanning law reflect strongly in the amounts and distribution of the data. For the astrometric data the distribution of scan directions was quite important, in particular for double stars. The scan direction is defined as the local angle between the reference great circle and the small circle of equal declination or latitude. The distributions of scan-directions as a function of ecliptic latitude are shown in Figure 14. The most important features are the missing scan directions near the ecliptic plane, and the large concentration of observations near ecliptic latitudes 47 . The asymmetry between scan directions near 0 and near 180 is the result of the motion of the z axis, which is moving along with the changes in the direction to the Sun when at southern latitudes, and moving against that direction when at northern latitudes (see also Figure 13). The peculiarities of the scan directions for every individual star reflect in the covariance matrix of the astrometric solution and have been preserved as such (see Section 8.7). The scanning law affected the photometric observations in two ways: the total number of available observations, and the statistics of the distribution of those observations. These aspects are shown in Figures 15, 16 and 17. The numbers of gaps and their lengths effectively limited the search for periodicity in variable stars to periods less than about 5 days, except for very frequently observed stars (see Section 9.4). The feature near the ecliptic plane in Figure 16 between 90 and 140 days results from the angular separation between stars near the ecliptic plane and the Sun during three to five months of the year. An additional, much weaker feature near 180 days is the result of a “closing error” in the scanning law that affected observations for a small number of stars close to the ecliptic plane (see also Figure 129). The satellite was made to follow the scanning law through the use of coldgas thrusters, which primarily corrected for the effects of torques acting on the satellite. The pointing was kept to within 10 arcmin from the nominal position, and was controlled by the on-board computer, using star mapper transit times for selected stars and gyro data. When the on-board attitude control was lost, which was usually the case after a perigee passage, the attitude parameters were reinitialized by the ground station. An elaborate software package was set up by A. McDonald at the mission control, ESOC, for recovering the pointing of the satellite from the continuous star-mapper data stream. This package operated during the last few months before the onset of the final breakdown of the satellite with such a success, that generally only very little time was lost after a perigee passage. When the two-gyro mode came into operation, this recovery speed was the main reason that some useful data were still obtained. 230 F. VAN LEEUWEN Figure 14. Distribution of the orientation angles of the scans for stars at different ecliptic latitudes. Note the “missing” angles for latitudes between 47 and the large number of scans parallel to the 47 . The peaks at 90 and 270 are the result of Sun-pointing mode ecliptic plane around observations. The full lines represent positive and the dotted lines negative latitudes. = THE HIPPARCOS MISSION 231 Figure 15. The total number of photometric observations per star as a function of ecliptic latitude. The bin size is 5 in latitude and 1 in number of field transits. The contour interval is 20 and the highest contour represents 200 stars per bin. Figure 16. Distribution of the lengths of gaps in the photometric data as a function of ecliptic latitude. The bin size is 5 in latitude and 1 in gap length. The contours rise non-linearly from 200 points to 8900 points per bin. 6. Mission operations time line The data quality and data return of the Hipparcos mission was affected by a wide range of intended changes and unintentional circumstances. The intended changes were generally aimed at improvements in quality and quantity of the data return: improved calibrations and improvements to the Input Catalogue. Various failures 232 F. VAN LEEUWEN Figure 17. Distribution of the number of gaps in the photometric data as a function of ecliptic latitude. Gaps longer than 1.5 days were counted. The bin size is 5 in latitude and 1 in number of gaps. The contours follow the relative distribution of gaps. The contour interval is 2 per cent, and the highest contours represent around 20 per cent of the data at the given latitude. and a few communication problems led to loss of observations, while a range of necessary calibrations and system tests also sometimes led to some temporary loss of usually very small amounts of data (gyro de-storage, refocusing and various routine instrument calibrations). Figure 18 shows schematically the main events that affected the data quality and return. Table V lists the various features shown in Figure 18. 6.1. ACTIVITIES LEADING TO IMPROVEMENTS OF THE DATA QUALITY After the failure of the apogee boost motor, it was essential to incorporate additional ground stations. Next to the main ground station in Odenwald (Germany), Perth in Australia and Kourou in French Guyana were incorporated at an early stage. Goldstone in the USA later took over from Kourou. Goldstone was positioned much better than Kourou to fill in the gap between Perth and Odenwald (see Figure 1). This brought the visibility of the satellite by ground stations to about 80 to 85 per cent. Odenwald alone could not have covered more than about 40 per cent. The incorporation of additional ground stations also implied that the number of satellite orbits without ground-station coverage at all was very small. Improvements in the data quality were obtained from ground-based calibrations of instrument parameters and Input Catalogue improvements. Most noticeable were the determination of the orientation of the grid assembly with respect to the scanning circle (by the First-Look processing in Utrecht) and the effect of improvements of star positions on the attitude reconstruction. The latter were initially obtained from THE HIPPARCOS MISSION 233 Table V Summary of the main events. Start and end times are given in days from 1989 Jan 0. TCE: Thermal control electronics; MDE: Mechanism drive electronics. Start 328 382 389 612 632 592 755 819 900 940 1001 1013 1161 1285 1285 1315 1453 End 596 1004 1163 1288 1396 1466 Code Event I1 I2 I3 T1 T2 a T3 V S H b I4 c T4 d e f Empirical correction of CCCM Analogue mode anomaly detected Grid rotation calibration implemented Payload thermal control anomaly Change from MDE1 to MDE2 Attitude lost, faulty RTAD initialization Change from TCE1 to TCE2 Spurious undervoltage Gas tanks swapped Gyro-3 heater failure Uplink command error Anomalous IDT voltages Data lost due to tape fault TCE2 anomaly, TCE1 fully broken down Non-z -gyro patch introduced Suspension of data acquisitions Gyro-2 anomaly, Sun-pointing for recovery accumulated star mapper positions as obtained at the RGO (see van Leeuwen et al., 1992a, Turon et al., 1992b and Section 8.3) and at later stages of the project from the preliminary sphere reconstructions. The improved positions led to better attitude control by the satellite and to improved pointing for individual objects. Further improvements for the attitude control came from the gyro orientation and drift calibrations at the RGO. 6.2. THE ROUTINE OPERATIONAL PHASE A major influence on the data quality and return came from the varying background levels. The main reason for this was the radiation background and the way this affected the star mapper photomultipliers. The star mapper background would normally increase during crossings of the Van Allen belts, but the amount of this increase fluctuated strongly. The Hipparcos mission operated during a time of high solar activity, and on several occasions the background levels for star mapper were so high, that attitude control by the satellite became impossible. This led in general to shorter data spans used in the great-circle reductions. Figure 19 shows the evolution of the background in the star mapper V channel over the mission at different orbital phases. The background levels are presented as the equivalent of 234 F. VAN LEEUWEN Figure 18. Summary of various events that affected the Hipparcos mission. Table V lists chronologically events that had direct bearing on the data reductions, and explains some of the symbols used in the figure. stellar signals of a given magnitude, to indicate to what magnitude stars could still be observed. The optimal focus of the telescope evolved slowly with time. In order to maintain a close to optimal focus setting, the focal plane assembly had to be shifted slightly every 2 to 3 weeks at the beginning of the mission, and at longer intervals later on. These refocusing events can be recognized in several calibration parameters, but should in most cases not affect the data, as they mostly took place outside the data aquisition time. An example of the effects of the refocussing is shown in Figure 20. The influence of some hardware failures on the focus of the telescope is clear from the changes in the modulation coefficients that can be associated with these events. 7. The data streams There were two primary and several secondary data streams. The primary data streams provided the compressed photon counts for the main detector (the image dissector tube or IDT) and the BT and VT star mapper detectors. These data formed THE HIPPARCOS MISSION 235 Figure 19. Star mapper background levels in the V channel at different orbital phases over the mission. The orbital phases range from -0.5 at perigee to 0.0 at apogee to 0.5 at perigee. The top graph also indicates events of high solar activity (just below the mag 5 level). 236 F. VAN LEEUWEN Figure 20. The effect of refocussing on the modulation coefficient M1 of the first harmonic in the I )C = 0:5 for the main-mission signal. The top graph shows the coefficient for stars with (V following field of view (FFOV) and the bottom graph for stars of the same colour index for the preceding field of view (PFOV). The discontinuities are the result of refocussing of the telescope. Various events that disturbed the thermal stability of the telescope are indicated in the top graph. The codes refer to Figure 18 and Table V. , part of a 23 kilo bits per second down link from the satellite. The secondary data streams provided information on gyro readings, temperatures, real time attitude determination, thruster firings and various status bits. All data was received and archived at ESOC in Darmstadt. At ESOC, the satellite telemetry was split into data streams and put on magnetic tapes for dispatch to the data reduction consortia. A Data Delivery Interface Document, defining the file structure, types and formats, had been established well before the start of the mission. The simulated software as prepared at the RGO for THE HIPPARCOS MISSION 237 the use in FAST, NDAC and TDAC had been written according to those specifications. When the real data came along, the reduction consortia were fully prepared and ready to receive it. Very little interpretation and no data decompression was applied to the data before being sent to the reduction consortia. The data collected over 1 week generally took 6 to 7 9-track 6250 bpi tapes for transfer. A total of around 1300 tapes each were sent over a period of almost 3 years to RGO (for NDAC), CNES (for FAST) and Tübingen and Heidelberg (for TDAC). After a period of testing of the interfaces and data reduction processes, which took place between the start of the first observations in November 1989 and the end of 1990, the production of all mission data on tape started in February 1991. The last tapes were received in December 1993. The first task in each consortium was the archiving of the data tapes. In FAST and TDAC the tapes were copied and catalogued for data contents, with the copies being used in the data reductions. In NDAC, the data was transferred from the tapes into ready-to-reduce data sets, which were stored on WORM-type optical discs, providing random access to all mission data. Of the data streams present on these tapes, four are of particular interest in the context of this review: The main detector or IDT data stream, containing the compressed IDT counts with some housekeeping giving information as to the association of counts with stellar images and details allowing the reconstruction of the pointing of the instantaneous field of view of the IDT (for NDAC and FAST); The star-mapper data stream, containing sets of 250 samples of compressed photon counts in BT and VT , centred on the expected transit times of reference stars to be used in the reconstruction of the satellite attitude (for NDAC and FAST); The Tycho data stream, compressed counts in BT and VT from the continuous stream of data obtained from the star mappers (for TDAC only); The satellite attitude data stream, containing details on the real time attitude determination by the on-board computer (this allowed the possibility to reconstruct the differences between the actual pointing of the satellite and the pointing which the on-board computer considered correct). This file also provided information on thruster firings and gyro readings. In addition there were files giving details on the orbit of the satellite, the housekeeping data (for example temperature measurements) and timings of eclipses of the Sun by the Earth or the Moon. The scientific data files are presented in more detail below, preceded by a description of the decompression of the photon counts. 7.1. DECOMPRESSION OF THE PHOTON COUNTS All photon counts were transmitted by the satellite to the ground station, and from the ground station to the reduction consortia, compressed in 1 byte (8 bits) integers. The compression law was semi-logarithmic, adding, when below saturation, a noise well below the intrinsic Poisson photon count noise. In the decompression, correc- 238 F. VAN LEEUWEN Figure 21. Verification of the implemented intensity transfer function (ITF) for the IDT data stream, using data for the brightest stars. The diagonal line is the expected relation (apart from an offset) for properly decompressed counts. Counts for the brightest stars, and in particular for Sirius at 1:5 mag, were still affected by uncorrected saturation. , tions for non-linearities of the photomultiplier responses were accommodated too, as were, to some extent, the effects of near saturation. Decompression therefore used tables rather than the mathematical form used for the compression of the counts by the on-board computer. The transformation to decompressed responses was described by the Intensity Transfer Functions (ITF). The low and intermediate intensity range of each ITF was well determined. The high intensity range, where the effects of saturation started to appear, was calibrated at the start of the mission, using mission data for very bright stars (see Figure 21). The IDT detector should have used for the 21 brightest stars (H p < 1:5) the analogue observing mode. Due to a synchronization error between the observing strategy and the switching between photon-counting and analogue mode, this mode could not be used. As a result, the data for the brightest stars suffer from saturation, affecting mainly their photometric results. The star-mapper data was compressed in the same way as the IDT data. Here, however, saturation affected the photon counts for stars of 2.5 mag and brighter. The saturation also affected the astrometric results for the brightest stars (see Figure 34), but in this case such results were all superseded by main-mission data. 7.2. THE IMAGE DISSECTOR TUBE DATA STREAM The Image Dissector Tube (IDT) data stream consisted of compressed photon counts obtained over 1/1200 s sampling periods. At nominal scanning speed (168.75 arcsec s,1 ), 8 such sampling periods roughly covered the crossing of THE HIPPARCOS MISSION 239 Figure 22. The average profile of the instantaneous field of view. Variations with star colour and with position on the grid existed. one slit (1.2074 arcsec) on the modulating grid. The sensitive area of the IDT, referred to as the instantaneous field of view (IFOV), could be moved very quickly across the grid, and had a diameter of about 30 arcsec (see Figure 22). It could obviously not distinguish between data from the two fields of view. Due to its small size, however, incidental images from the other field of view appearing within the IFOV at the target position were relatively rare. Their influence depended very much on the magnitude difference between the target object and the incidental object. Significant disturbances were therefore more common for faint stars than for bright ones. The sampling periods were grouped in such a way that near simultaneous observations were obtained for up to 10 stars over a period of an observational frame, 32/15 s. The basic unit of these observations was the “slot”, a period of 1/150 s covering 8 samplings and being the minimum number of consecutive samplings for the observation of a star. Sets of 20 slots, forming interlacing periods of 2/15 s, were distributed among up to 10 programme stars visible in the field of view. The number of slots assigned to each star depended on the magnitude and on the competition from other stars (see also Figure 48). The observational frame was made up of 16 interlacing periods, and was the unit time interval for the IDT data reductions. Between the 16 interlacing periods, up to three different distributions of observing time could be implemented. This way it was possible to accommodate one star entering and one star leaving the field of view during the time-span of an observational frame. Such observations were referred to as “partially observed”. Only brighter stars were ever partially observed. Fainter stars were only “fully observed”, i.e. when visible during the entire observational frame. A reversal of priorities between stars from one frame to the next helped to establish a proper overall distribution of observing time. This often meant that the 240 F. VAN LEEUWEN faintest stars, requiring many slots to get reliable observations, were observed in alternating frames only. The repositioning of the IFOV (moving from one stellar image to the next) required a small but not negligible amount of time. The first sampling period after a repositioning was therefore not used in the data reductions. The positioning of the IFOV was controlled through coil currents, for which the values were included in the data stream. The coil currents were calibrated at regular intervals; these calibrations were provided in the form of coil current calibration matrices (CCCM) on the data tapes. They allowed a reconstruction of the pointing to be made, and a comparison with the a posteriori reconstruction of the best pointing. Figure 23 shows the evolution of the pointing accuracy over the mission. The on-board computer determined the amounts of observing time to be spent on each object, using a file with stellar positions and priorities as provided by the ground station, and the on-board attitude reconstruction to see which objects would be visible and where these would be situated in the field of view. The result of the observing strategy was that a variable amount of observing time was assigned to the observations of each star, which reflects in the accuracies of the reduced data (see Figure 48). The modulating grid produced a signal consisting of a zero level (signal plus background) and a first and second harmonic sinusoidal modulation. The astrometric data were derived from the modulation phases; the mean intensity and modulation amplitudes provided the photometric data. The relations between the modulation amplitudes and between the phases of the first and second harmonics for single stars were a function of stellar colour and position in the field of view. Such relations, described by a calibrated optical transfer function (OTF, see Section 8.4.5), were used to recognize the superposition of more than one image (both as a result of double stars and as a result of accidental superpositions), in which situation these relations no longer applied (see Section 8.4). The photometric passband for the IDT detector, referred to as Hp, was designed for optimal return on astrometric data. It was therefore a broad band, which extended towards the red, but with an effective wavelength close to the Johnson VJ band. Figure 24 shows the Hp passband in comparison with passbands in the Johnson and Cousins systems. The count-rate for the IDT detector depended on the stellar magnitude, colour and time of the mission. The detector response decreased over the mission as a result of radiation darkening of the relay optics in the IDT. The effect was most pronounced for shorter wavelengths, and caused changes in the passband (see Section 9.3). For an 8th magnitude star with B , V = 0, the mean response was almost 7000 counts s,1 at the beginning of the mission, dropping to 4000 at the end; the relative drop for red stars was much less (see Figure 84). THE HIPPARCOS MISSION 241 Figure 23. The observed pointing accuracy of the IFOV over the mission. The graphs show the fraction of observations for which the reconstructed IFOV pointing was within the indicated radius from the actual position of the stellar image. Only observations within 9.5 arcsec were used, those outside were considered to be associated with failures of the on-board attitude control loop. The overall improvements in the satellite attitude, the Input Catalogue and the grid geometry calibration led to the increase in the pointing accuracy over the mission. The last part of the mission was hampered by bad attitude due to gyro failures. 242 F. VAN LEEUWEN Figure 24. Normalized response curves for the Hipparcos on (from left top right) the BJ , VJ , RC and IC passbands. Hp passband (solid line), superimposed Figure 25. The BT and VT passbands (solid lines) compared with the BJ and VJ passbands (dotted lines). 7.3. THE STAR-MAPPER DATA The star-mapper data stream was obtained from the light detected through any of the four slits of both slit groups, and from both fields of view. The slit-width was 0.9 arcsec and the span of the slits perpendicular to the scan direction was 40 arcmin. The slit spacings were chosen such that at nominal scanning speed (168.75 arcsec s,1 ) the slits were encountered at intervals of 40, 60 and 20 sampling periods of 1/600 s. At the nominal scanning speed one sampling period was equivalent to a movement of 0.28 arcsec. The light behind the slits was split into THE HIPPARCOS MISSION 243 Figure 26. Photomultiplier tube signal (blue channel) for the transit of one star over vertical slit group. The sampling period is 600 Hz. The four peaks, corresponding to the four aperiodically spaced star mapper slits, yield the star intensity and transit time. The observation is for a transit of a BT 6:62 mag star (HIP 35946). = two passbands, referred to as BT and VT , more or less resembling the B and V passbands in the Johnson system (see Figure 25). The star-mapper data were grouped in sets of 25 sampling periods, covering 1/24 s, equivalent to one telemetry frame. The data of 256 of such groups covered one telemetry format (32/3 s), equal to 5 observational frames of 32/15 s. The Tycho data reductions received all the star-mapper data in units of telemetry formats. NDAC and FAST only received a selection of data, associated with stars bright enough to be of use in the on-ground attitude reconstruction process. Centred on the expected transit times of such stars (as based on the real time attitude determination, see the next section), 10 groups of 25 samples were extracted. In cases when the on-board attitude was incorrect and not converged, this meant that the extracted samples would partly or completely miss the transit signal. Extracted signals were identified in the data stream through the star number, the slit group and the field of view for the transit. Figure 26 shows an example of a decompressed star mapper transit signal for a relatively bright star. The average number of transits extracted this way was 15 per minute, originating from both slit groups and both fields of view. In the Tycho data stream the record was identified with a format counter and the ground-station time-tag, but identification of transits was left to the TDAC consortium. The average number of transits present in one format of star mapper data (32/3 s) down to the limiting magnitude 11, was 33, but there were considerable variations. 244 F. VAN LEEUWEN 7.4. THE ON-BOARD ATTITUDE DATA The on-board attitude information was provided in the form of three streams: the gyro measurements, the thruster firing information and the results of the on-board real-time attitude determination. The knowledge of the thruster firing timings, and to some extent their strengths, was crucial in the data reductions. Thruster firings caused discontinuities in the rotation rates for the satellite axes which had to be accounted for in the reconstruction of the satellite attitude. By calibrating the changes in the inertial rates around the satellite axes with the lengths of the thruster firings, constraints could be imposed on modelling of the attitude across firings. In addition, the calibration results provided further insights in the working of the spacecraft. Figure 27 shows the results of such calibrations over the entire mission. Various features can be distinguished: the discontinuity at day 900 is due to the change of cold-gas tank, when tank 1 was almost empty; the slope in the relations is related to the depletion of the gas-tanks, with the decrease in pressure being over-compensated; some features are due to temperature variations in the spacecraft and can be attributed to periods of very long eclipses. Here too, the decreased pressure caused by the lower temperature was over-compensated and created a higher thrust; the variability of the thrusts was about twice as high for the second tank, possibly related to its positioning on the spacecraft deck; the results of thrusts could be predicted to within 0.02 arcsec s,1 . Thrusters were fired only at the start of an observational frame. The main mission data collected during such a frame was discarded. Discontinuities in the rotation rates were also occasionally caused by the satellite being hit by micrometeorites. Such discontinuities could only be recognized from an examination of the gyro data (before the start of the data reductions, by NDAC) or on the basis of the badly fitting attitude reconstruction results (by FAST and NDAC). Further problems occurred during the monthly gyro de-storage exercises (see Figure 28), which created sudden temporary torques within the satellite while a redundant gyro was spinning up. Also these situations could only be recognized from the gyro data, as they were not recorded in the telemetry. The real time attitude determination data was a record of what the satellite considered to be the pointing of its three principle axes and the inertial rotation rates around these axes. These data were provided at intervals of 16/15 seconds and were very useful for checking, as part of the data reductions, the convergence and stability of the on-board attitude reconstruction, and thus of the quality and reliability of the scientific data. The most direct comparison that could be made was with the observed star mapper transit times (see Section 8.1 and Figure 29). The real time attitude determination on-board the satellite was based on transit times for selected stars and gyro data. The transit times were used to determine THE HIPPARCOS MISSION 245 Figure 27. The evolution of the thrusts created by the cold-gas thrusters. Thrusters causing positive velocity changes are indicated with filled symbols, those producing negative changes with open symbols. The discontinuity is caused by a change of gas tank, the drift is caused by the emptying of the gas tanks and a slight over-compensation for loss of pressure. pointing directions: for transits through the vertical slits they provided the scan phase estimate, and for the transits through the inclined slits of the two fields of view the position of the satellite’s spin axis. Rotation rates were derived in the same determination from the observed changes in positions. These rates were 246 F. VAN LEEUWEN Figure 28. An example of the influence on the rate of change of the inertial rates caused by a gyro de-storage exercise. The readings of gyro 4 are shown, calibrated to rates relative to the nominal scan velocity. The redundant gyro 3 is spun up at position a. Its spin axis is aligned with the input axis of gyro 4 (parallel to the satellite’s z axis), which is noted in the accelerations. At position b the redundant gyro 3 is spun down, causing a deceleration. The discontinuities in the rates are the result of thruster firings, which also compensated the changes in the satellite’s angular momentum that resulted from the de-storage exercise. used to calibrate on-board drifts for the gyros that were used to bridge the gaps between accepted star mapper transits. The angles and rates of the real time attitude determination could be overwritten through commands from the ground station. The updating mechanism was regulated by a Kalman filter. 8. Overview of the main data reductions Many features of the Hipparcos data reductions were first designed and proposed by L. Lindegren, most noticeably, already in 1976, the division of the data reductions from the phase estimates to the astrometric parameters into three discrete processes. This so-called three-step method for the data reductions was used by both data reduction consortia to extract from the individual phase estimates the final catalogue of astrometric parameters and was crucial in the feasibility of the project. The three-step method split the data reductions into the processes of the great-circle reductions, the sphere reconstruction and the astrometric parameter determination. As data reduction packages were developed and tested, the divisions shifted slightly, and other aspects were also incorporated: the initial processes of the IDT and star mapper photon count reductions, and the attitude reconstruction. At a later THE HIPPARCOS MISSION 247 Figure 29. The differences expressed in 600 Hz sampling periods between observed transit times and real time attitude determination (RTAD) based predictions. One field of view is converged from the start, the second field of view converges around frame 2400, after which there is general agreement between the observed transit times and the RTAD. Transits through the vertical slits are indicated with squares, for the inclined slits with triangles for the upper branch and stars for the lower branch. Filled symbols refer to the preceding field of view, open symbols to the following field of view. The dotted vertical lines indicate thruster firings. stage in the data reductions, when merging of the FAST and NDAC data was investigated, it was realized that a determination of the final astrometric parameters should be done on the combined results of the great-circle reductions and sphere reconstruction from FAST and NDAC, rather than as weighted averages for the independent determinations by the two groups. Thus, the determination of the astrometric parameters is described in the context of the data merging, though preliminary, consortia based, determinations took place too. Extensive descriptions of the final data reductions can be found in Volumes 3 and 4 of ESA (1997) for the Hipparcos and Tycho reductions respectively, and summaries by Kovalevsky et al. (1992), Lindegren et al. (1992c) for the main-mission data, and by Høg et al. (1992a) for the Tycho data. This section describes the reduction steps more or less in the sequence in which they took place. The following is a brief summary of these processes, the sections in which they are described and the way they are interlinked. A: Star mapper reductions. Described in Section 8.1. Used raw star mapper counts and scan velocity estimates from the first step in the attitude reconstruction (B ). Provided the transit data for the final step in the attitude reconstruction. The reduced data were also used in catalogue updating (C ) and star mapper photometry (H). B: Attitude reconstruction. Described in Section 8.2. Used star mapper transit times (A) and auxiliary information from the satellite (gyro readings, thruster 248 F. VAN LEEUWEN firing information). The reduced data were used in the reduction of the image dissector tube data (D ) and the great-circle reductions (E ). C : Catalogue updating. Described in Section 8.3. Used star mapper transit time residuals obtained from attitude reconstruction (B ), and at a later stage early sphere solution results. Also collected photometric data from the star mapper and the main detector. Used in star mapper reductions, attitude reconstruction, great-circle reductions and sphere solution, as well as in photometric reductions and double star analysis. D : Image dissector tube (IDT) data reductions. Described in Section 8.4. Used raw photon counts from main-mission detector and attitude reconstruction results to relate measurements to proper positions on the modulating grid. Derived the modulation parameters: mean intensity, phases and modulation amplitudes. Modulation phase information was used in the great-circle reductions (E ) and mean level and amplitude information used in the main-mission photometry. Full modulation information was also used in double star analysis. E : Great-circle reduction. Described in Section 8.5. Used phase information from IDT reductions (D ) and the star mapper based attitude reconstruction (B). Derived abscissae for stars along reference great circles and an improvement of the along-scan attitude. The abscissae were further processed in the sphere solution (F ), the improved attitude was used in the Tycho reductions (H) and in the double and multiple star analysis. F : Sphere solution. Described in Section 8.6. Used data from great-circle reductions (E ). Determined zero-points of reference great circles and preliminary astrometric parameters. Input for the catalogue merging process (G ). G : Merging and astrometric parameters. Described in Section 8.7. Used data from FAST and NDAC sphere solutions (F ), derived correlations and orientation corrections, and determined the final astrometric parameters. Includes the transformation of the final catalogue to the ICRS. H: Tycho reductions. Described in Section 8.8. Used the raw star mapper photon counts together with main-mission attitude reconstruction results, based on the main mission astrometric catalogue. Produced a final catalogue aligned with the ICRS, containing photometry and astrometry for just over one million stars. 8.1. STAR MAPPER REDUCTIONS, INCLUDING TYCHO In order to extract from a star mapper signal the transit time and intensity for a star, several steps had to be taken, involving calibrations, recognition procedures (including the recognition of any disturbing signals) and signal fitting. Many of the calibrations were part of iterative processes, where preliminary calibrations were used in the data processing first, followed by the first actual calibrations and implementation of the first calibration results in a new reduction. These steps were often repeated two to three times before calibrations were accepted as final. Thus, THE HIPPARCOS MISSION 249 in the description of the calibrations, results from the transit time and intensity determinations are used, and similarly in the description of the transit time and intensity determinations the results of calibrations have been used. The calibrations of the grid geometrical and response characteristics are presented first, followed by aspects of the actual data reductions. 8.1.1. Single slit response functions The single slit response functions described the normalized detector responses to the crossing of a single slit by a point source, integrated over the 1/600 s periods, and described as functions of the distance to the slit centre. These response functions were obtained from accumulated transit data for stars brighter than 8th magnitude and with a low background count. In the data reductions, transit times and signal intensities had been determined for these stars, as well as background levels and local scan velocities (see Section 8.1.6). Given this information, the observed counts for all four slits could be related to a distance from each assumed slit centre, and scaled (or weighted) according to the measured signal intensity. These values were binned according to distance from the slit centre (at a time resolution of 9600 Hz by NDAC and 5400 Hz by FAST), slit, slit group, within each slit group the upper or lower branch, channel (BT or VT ), and field of view. Data for 20 000 to 30 000 transits, covering a time span of 2 to 3 months, were collected for every single calibration. This allowed a check on any significant changes in the response functions over the mission. Only small changes in the wings of the response functions were noted. The accumulated weighted data in each bin were divided by the accumulated weights to provide response measurements. These response measurements were fitted using cubic-spline functions Rq in least squares solutions. The index q in Rq represented the 16 different combinations of slit group (the vertical slits and the inclined slits), branch, field of view and passband. The centre t0 of each curve Rq (t , t0 ) was defined by the point halfway responses of Rq = 0:5 on the rising and descending branches. The curves were normalized to 1.0 at peak (NDAC) or at t = t0 (FAST). The derivatives of the response curves were determined directly from the cubic-spline fits. These derivatives were used in the determination of transit times (Section 8.1.6). The widths of the response functions at a response level of 0.5 ranged from 0.89 to 0.90 arcsec for the vertical slits, and from 1.00 to 1.07 arcsec for the inclined slits. The response curves showed significant differences between various combinations of slit group, branch, passband and field of view. Figure 30 shows two examples of reconstructed single slit response functions and the corresponding derivatives. No significant differences in response functions were detected as a function of colour index, but no sufficient accumulation of data was possible for the extreme red stars (V , I > 2). The measurements for stars brighter than mag 2.5 were affected by saturation of the highest counts in the signal. Here the knowledge of the wings of the response 250 F. VAN LEEUWEN Figure 30. Reconstructed single slit response functions (left) and their derivatives (right) for the inclined slits (solid line) and vertical slits (dotted line), upper branch, in the preceding field of view, recorded in the BT channel. The derivative is given as the change of response per 600 Hz sampling period at nominal scan-velocity, when one sampling period corresponds to 0.28 arcsec. Table VI Calibrated distances between the slit positions dj in the star mapper vertical slit group. Interval Arcsec mm ground (mm) d1 to d2 d2 to d3 d3 to d4 11.247 16.867 5.623 0.07638 0.11455 0.03819 0.07640 0.11460 0.03820 functions became very important. In TDAC special calibrations of these wings were carried out, also in order to be able to determine proper background levels in the presence of very bright stars. 8.1.2. Grid geometry calibrations There were four aspects of the grid geometry that required calibrating: the spacings between the individual slits; the position of the inclined slit group with respect to the vertical slit group; the orientation of the slit groups; and the medium-scale distortions for each group. The spacings of the slits were very accurately calibrated as part of the above described single slit response function calibrations. By determining the responses for each of the four slits individually, the corrections to the assumed spacings were determined by cross-correlating neighbouring responses. The values that were THE HIPPARCOS MISSION 251 obtained were very close to the ground-based grid measurements, as can be seen from Table VI. The orientation of the slit groups and their relative positions were calibrated as part of the attitude reconstruction, either direct as instrument parameters or a posteriori from the residuals of the star mapper transit times after fitting of the satellite attitude. These calibrations were also carried out as part of the Tycho data processing, from which a record of star mapper related instrument parameters as a function of time over the length of the mission was obtained (Høg et al., 1995). The medium-scale distortions described small deviations of the grids from linearity. These distortions had been determined from measurements of the grid in the laboratory before the start of the mission. They ranged from 0.01 to 0.05 arcsec when projected on the sky. Their proper implementation was checked using the star mapper transit time residuals obtained from the attitude reconstruction process. Empirical calibrations of the medium-scale distortions were obtained in the Tycho astrometric reduction. 8.1.3. The star mapper transit signal The star mapper photon-counts Nk;c , collected at times tk by either channel c = B; V during the passage of a star, were described as a sequence of statistically independent Poisson-distributed counts, having the following time-varying average: E[Nk;c ] Ik;c XR t , 4 = Ib;c + Is;c j =1 q ( k j ); (2) where Ib;c and Is;c are the intensities of the background and star respectively. Rq is the single slit response function as defined above and j are the transit times for the four slits. The transit time for the slit group was defined as the unweighted mean = 0:25 4j =1 j . The transit times j were linked to the individual slit positions dj through the instantaneous local scan velocity vq : P j = + dj =vq : (3) In the NDAC applications the scan velocity vq was obtained from the gyro data, and a single mean transit time estimate was obtained. In the FAST determinations the transit time per slit was obtained, and from this the mean transit time . 8.1.4. Transit detection The transit detection processes that were developed in the different consortia all had one aim: to reduce the signal of four peaks to a signal of one peak, in order to be able to distinguish more easily the often partly overlapping signals of different transits. This involved various filters and algorithms to examine the filtered results. The task of a filter was to recognize the characteristic pattern of the four slits. Filters were generally applied to the combined BT and VT photon counts for increased 252 F. VAN LEEUWEN Figure 31. NDAC processing of star-mapper data for a transit of the double star HIP 51560/51561 through the vertical slits. The first two bars show the original counts, the third bar the cumulated signal J , and the fourth bar the filtered signal F . Between the V and B signals is indicated which samples were excluded from the calculation of the background. Below F are indicated the recognized transits: here the program star transit is fainter than the companion. The arrows below indicate the predicted position, as based on the on-board attitude. The lower half of the graph shows the original B counts folded with the final signal fit, which excluded the first 26 and last 35 samples, within which it was not possible to identify transits unambiguously. signal sizes. The filter as implemented in the reductions for NDAC at the RGO is given here as an example. The FAST filter can be found in Section 6.4 of Volume 3 and is also partly described in the context of the FAST background determination in Section 8.1.5. The TDAC non-linear 4-peak filter is described in Chapter 2 of Volume 4 of ESA (1997). A multiplicative filter was implemented by NDAC. It was applied to the the following reduced resolution signal: X (Ik,1;c + Ik;c , 2I~b;c ); J (k=2) = (4) k = 2; 4; 6; : : : 250: = c V ;B All negative values of J (k=2) (counts below the estimated background level I~b;c ) were set to zero. This signal was subjected to the multiplicative filter, which consisted of four Dirac functions, spaced like the expected slit responses: 20, 30 and 10 intervals at 300 Hz: 1=4 Y J (l + n) F (l ) = ; n = ,30; ,10; 20; 30: (5) n THE HIPPARCOS MISSION 253 Special provisions were made for signals with only two or three out of four peaks visible within the 250 counts provided. A transit signal was recognized from a cluster of F (l) values different from zero. More than one cluster could be recognized, and up to four could be accounted for in the further processing. The peak values of the clusters provided the first estimates of the transit times and indicated to the further processing which samples were associated with star transits, and which samples could be used for background determination. With more than one transit present, the one closest to the expected brightness and transit time was selected as the intended transit (the transit of the star for which the counts were extracted). For known double stars the expected transit time difference between the components was incorporated in the recognition mechanism. Figure 31 shows these steps in an application to the star mapper transit signal for a known double star. In the Tycho data stream the identification of transits with detections was a different problem. A special process, referred to as ‘prediction’ produced initially predicted transit times based on the Tycho Input Catalogue (Egret et al., 1992) and the real time attitude determination. At later stages the NDAC attitude reconstruction results were used in combination with the revised Tycho Input Catalogue (Bässgen et al., 1992, Halbwachs et al., 1992, see also Section 8.8). 8.1.5. Background determination The background determination was important for the fitting of the single slit response functions, on which both the transit time and the signal intensity determinations depended. Two examples of the background determinations are presented here, the implementations by FAST and TDAC. In the FAST processing a set of five variables Tj (m); j = 0; 1; : : : 4 was obtained from the convolution of the cumulated counts in the B and V channels with the single slit response function: Tj (m) = X X Nk;cRq tk , tm , dj =vq ; ( c=B;V k ) (6) where dj = 0 for j = 0, and defined in Table VI for j 6= 0. This signal, shown in Figure 32, was transformed in a set of Boolean test variables Lj (m), which gave a value of one when Tj (m) was above a suitably chosen test value s, indicating the presence of a star transit signal, and zero otherwise. A generic sample m was accepted as a background sample if L0 (m) = 0 and it was separated by at least n0 samples from a position for which L0 (n) = 1 applied. The last condition avoided the inclusion of samples in the background calculation that were affected by the wing responses of a single slit transit. The background samples detected in a record were averaged to produce the (local) background estimates for the transit: I^b;VT = 1 X Nk;V ; n k2A T I^b;BT = 1 X Nk;B n k2A T (7) 254 F. VAN LEEUWEN Figure 32. Top: a photon count record for a 6.6 mag star (HIP 47115) through the vertical slit group, with the estimated average signal superimposed. The abscissae are in sample units (600 Hz), the ordinates are in counts per sample. A spike, which occurred in the V channel, is indicated by an arrow. Bottom: the corresponding convolution T0 (m) as provided by the FAST processing. where n is the number of accepted background samples and A is their sequence in the record. A running estimate of the background, Ib , was maintained during a data set, based on the estimates obtained from successive semi-contiguous records. The running estimates were updated at each transit i using the local estimates I^b;B (i) and I^b;V (i). A first order Kalman filter was used for each channel: Ib;c (i) = [1 , K(i)]I b;c (i , 1) + K(i)I^ b;c (i) (8) where the time varying gain was set to K(i) = n(i)=(n(i) + 250), depending on the number n(i) of detected background samples, but restricted to the range THE HIPPARCOS MISSION 255 Figure 33. The estimated background intensity over an orbit on 10 June 1991, coinciding with a period of very high solar activity. The background levels detected at apogee (centre of the graph) were more than 10 times higher than under normal conditions. The top curve represents the V channel, the lower curve the B channel. The Van Allen Belt crossings are indicated by VA, resets of the Kalman filter by arrows. 0:05 K (i) 0:5. This range was designed to trade off between two contrasting goals: attenuating the photon noise error (typically when the background was low and few samples could be detected), and tracking the real background variations. The Kalman filter was reset when there was a significant time gap in the data. An example of the running background estimate for a time interval with particularly strong background variations is shown in Figure 33. In the Tycho processing two kinds of background estimates were obtained, a global background based on the median of all 6400 samples contained in a telemetry format, and a local background, based on 192 samples around every estimated transit position. The latter was used when background was disturbed due to very bright stars, dense star clusters, Galactic plane, zodiacal light, radiation or stray light (Wicenec & Bässgen, 1992). With its sensitivity to the radiation from outside the satellite, the star mapper background provided a kind of record of the variations in the radiation environment to which the satellite was subjected, both within each orbit and over the mission (see also Figure 19). This allowed an investigation into the evolution of the activity of the radiation-belts over the lifetime of the mission (Daly et al., 1994). The star mapper background records also provided information on the zodiacal light and the glow of the Galaxy (Wicenec & van Leeuwen, 1995). A complete record of the star mapper background over the mission is kept at the RGO. 256 F. VAN LEEUWEN 8.1.6. Transit time and signal intensity determination All methods used to determine transit times aimed at a way to optimize the fitting of the single slit response functions to the data. Here the implementation by NDAC is shown as an example because of its capability of handling more than one detected transit at a time, and the way this assisted some of the main-mission double star processing. The estimates of the transit times and signal intensities were done as two independent solutions in an iterative loop. All estimates were obtained using linear least squares. A first solution was obtained using unit weights for all observations, subsequent solutions used weights according to the estimated signal plus background for each sample. This was effectively equivalent to a joint maximum likelihood estimation for the transit times and the intensities. The basic equation for the intensity determinations is Equation 2, which, rewritten in the form of single transit times for m transits, gives: E[Nk;c ] Ik;c = Ib;c + m X s=1 Is;c 4 X j =1 Rq (tk , s , dj =vq ); (9) The equation for the transit time corrections is obtained by differentiating Equation 9 with respect to the transit times s . Information from the BT and VT channels was used together to determine single transit time corrections ds : m X X c=B;V s=1 dIk;c = m X X c=B;V s=1 q Is;c @R @ ds: (10) s This is where the derivatives of the single slit response functions, introduced in Section 8.1.1, are used. These derivatives are defined as ,Rq0 = @Rq =@ . The values of dIk;c are reflected in the differences between the observed counts and the predicted counts using the previous estimates of the transit times ~s and signal intensities I~s;c . Thus, we can write: m X X c=B;V s=1 X c=B;V ds , I~s;c Nk;c , Ib;c , m X s=1 4 X j =1 Is;c ~ Rq (tk , ~s , dj =vq ) 0 4 X j =1 = Rq (tk , ~s , dj =vq ) : (11) A practical limit of m 4 signals was applied, beyond which the resolution of individual signals became too bad. In order to provide the proper Rq and vq for all identified transits, each would have had to be identified with a field of view and a slit group. This was considered too time-consuming when the data reductions were performed and of little effect to the quality of the parameter estimates of the target transit. Thus, all parasitic transits were treated with the scan velocity and slit responses of the target transit. THE HIPPARCOS MISSION 257 Figure 34. The formal standard errors ( , in arcsec) of the transit time determinations as a function of the total signal intensity, transformed to an approximate magnitude scale. The graph shows results for 9400 reduced star mapper transits obtained over a period of 9 hours in January 1990. The line represents the development of accuracies when only affected by the photon noise of the signal. An example of formal standard errors of the transit time determinations is shown in Figure 34. There were on average some 900 star mapper transits of Hipparcos stars available for every hour of observations, varying from 600 to 1200 depending on the inclination of the scanning circle with respect to the Galactic plane. Of these transits, half contributed to the determination of the scan phase, and a quarter each to the determination of the two coordinates of the spin axis of the satellite (see Section 8.2). The average standard error on a transit was approximately 0.03 arcsec, with a range from 0.1 to 0.003 arcsec. Transits with transit-time accuracies worse than 0.1 arcsec were rejected, as well as transits failing basic statistical tests for the model fits. In the Tycho astrometric processing (Section 8.8), an average of about 10000 transit times were derived from every hour of observations. These transit times were transformed into one-dimensional positions through the reconstructed satellite attitude. 8.1.7. Duplicity effects Double stars with separations above 1.5 to 2 arcsec could be resolved as separate transits in the star-mapper data processing. Below these separations a merged signal was observed and processed. The reconstructed intensity of such a signal was a function of not only the intensities of the component stars, but also of their projected separation. Thus, unresolved double stars can show up in the Tycho photometry file with magnitudes that are a function of the scan-position angle. The results of a search for such cases among 480 000 Tycho stars with VT < 10:5 are indicated in the Tycho Catalogue (field T49). For 8441 systems clear indications of duplicity were found (indicated by ‘D’ in field T49 in the Tycho Catalogue, ESA, 1997). 258 F. VAN LEEUWEN Figure 35. Distribution of the ordinates for all measurements used in the great-circle reductions, relative to their reference great circles. The solid curve represents the NDAC data, the dotted curve the FAST data. The Tycho Epoch Photometry Annex file (tyc ep.dat on disk 4) provides magnitudes and scan-position angles (field TT9) for a selection of 34 446 stars from the Tycho catalogue. 8.2. ATTITUDE RECONSTRUCTION This section describes the satellite attitude (the rotations of the satellite around its major axes and the resulting pointing of the satellite as a function of time) and the processes to reconstruct it using star mapper transit times, star positions, gyro data and torque modelling. The emphasis here is on the modelling of the motions and the information that was extracted from it. The Hipparcos data have provided an unique insight in the torques working on a satellite, with many of the details still to be resolved. Much more detailed accounts of the methods used to reconstruct the attitude are presented by Donati et al. (1992), van Leeuwen et al. (1992b) and Chapter 7 of Volume 3 of ESA (1997). This section is divided into four parts, describing the aims and target accuracy, the dynamical model, the FAST and NDAC approaches to the attitude reconstruction and the final accuracy assessment. The dynamical model and the NDAC processing were closely interlinked: the modelling used the reconstruction results, and the reconstruction used the model as a first approximation. The FAST processing was, to some extent, independent of this model. THE HIPPARCOS MISSION 259 8.2.1. The aims and target accuracy The attitude reconstruction for the Hipparcos satellite required an accuracy well beyond any preceding, and so far also following, satellite. The accuracy with which the pointings of the three main axes of the telescope reference frame (x, y, z) were to be determined was mainly set by the requirements of the great-circle reduction process: here transit positions were projected onto a reference great circle, to be related to other transits. The maximum projection distance (the ordinate of a star in the great-circle reference frame) was 1:5 . Figure 35 shows the histograms of all observed ordinates over the mission by FAST and NDAC. The small difference between the two curves resulted from the slightly higher tolerances of the NDAC attitude reconstruction to the more extreme observing conditions close to perigee, thus giving generally longer stretches of data per reference great circle (see also Figure 49). For 99.9 per cent of the data the ordinates were less than 1.278 . The target accuracy for the attitude reconstruction set before the start of the mission was for the root mean square of the errors to be less than 0.1 arcsec. For an ordinate at 1 this would add a noise of 1.6 mas in the form of small projection errors to the IDT transit time estimates, which is well below the noise on the phase estimates for most individual IDT transits (see Figure 48). 8.2.2. The satellite dynamical model, I: The equations of motion A first attempt to reconstruct and model the torques working on the Hipparcos satellite was presented by van Leeuwen et al. (1992b) on the basis of a small amount of mission data used for testing data interface and reduction software. The dynamical model of the motions of the satellite presented here describes the motions around its major axes, and not the orbital disturbances. The model is based on a mixture of observational evidence and theoretical considerations, and will be illustrated by the observations of various torques that affected the satellite. A reconstruction of torques was possible for the Hipparcos satellite due to the very high requirements put on the attitude reconstruction, and the fact that even these high requirements were most of the time surpassed by a significant margin. This allowed the resulting attitude to be examined in the form of rates and accelerations to a very high precision. The attitude of the satellite is described with respect to an orthogonal set of three axes, (x, y, z), fixed with respect to the satellite and defined by the geometry of the star mapper grid: the xy-plane was defined by the great circle through the corner points of the inclined slits as seen through the preceding and following fields of view. The x axis was defined halfway these two points, the z axis perpendicular to the plane, and the y axis completed the triad. This is schematically shown in Figure 36. As a result of a small rotation (5 arcmin) of the grid assembly, the centre of the main grid did not coincide with the xy plane. The satellite is assumed to be a rigid body, which, considering the very small torques working on it and the rigidity of the satellite that was required for the proper 260 F. VAN LEEUWEN Figure 36. Definition of the major axes with respect to the telescope and the grid geometry. The xy-plane is defined by the great circle through the corner points of the inclined star mapper slits as seen through the preceding and following fields of view. operation of the instrument, seems a reasonable assumption. Thus, its motions could be described by the Euler equation: I d! dt =N , ! I!; (12) where I is the inertia tensor of the satellite, N the external torques, and ! the inertial rotation rates around the axes of the satellite reference frame. An approximate analytical solution for the Euler equations was obtained by Bois (1986) in an application to a model of Hipparcos affected by solar radiation torques. In the analysis of the satellite attitude both the rotation rates ! and the accelerations dd!t were measured. Following the approximations made in Chapter 8 of Volume 3 of ESA (1997), Equation 12 can be written as: 0 N =I 1 0 ! ! (I , I )=I , ! ! I =I 1 yy xx x z xy xx d! @ x xx A @ y z zz Ny =Iyy + !z !x(Ixx , Izz )=Iyy + !y !z Ixy =Iyy A : dt Nz =Izz !x!z Iyz =Izz + !y !z Ixz =Izz (13) This equation, expressed in the observable inertial rates and accelerations, allowed for the calibrations of the inertia tensor element ratios as a function of time over the mission. Changes in the inertia tensor took place primarily as a result of the depletion of the cold-gas tanks used for thruster firings. Table VII gives the calibrated inertia tensor elements scaled to a fixed, ground-based value of Ixx . This calibrated inertia tensor was subsequently used for all other torque calibrations, realizing, however, the arbitrary scaling to the Ixx element, which may be responsible for some small drifts in the observed torques. In most cases, however, observed torques were a function of the same inertia tensor element ratios as those calibrated above. THE HIPPARCOS MISSION 261 Figure 37. Removal of the Euler-equation cross-terms and the gravity gradient contributions from the observed accelerations. Top graph: the observed accelerations around the x axis (black) and the predicted gravity gradient contribution (grey). Bottom graph: the remaining accelerations. The data cover one orbit (orbit number 307, March 19, 1990), with apogee at phase 0.5. The effect of a solar eclipse is clearly visible between phases 0.17 and 0.32. Discontinuities in the top graph relate to thruster firings, causing discrete changes in the rotation rates. 262 F. VAN LEEUWEN Table VII The inertia tensor elements in kg m2 for the Hipparcos satellite. The reference epoch is day 70 in 1991 and t is measured in years from that epoch. The inertia tensor has been scaled to a fixed value of Ixx . x x y z 534:220 3:459 2:031 y + 0 000 , 0 469 + 0 082 : t : t : t 3:459 590:680 0:806 , z , 0 469 + 0 209 + 0 005 : t : t : t 2 031 + 0 082 ,0 806 + 0 005 459 420 , 1 390 : : t : : t : : t 8.2.3. The satellite dynamical model, II: The gravity gradient After the determination of the inertia tensor of the satellite, the torques due to the gradient of the Earth’s gravitational field across the satellite were fully described. Following Spence (1978), the gravity gradient induced torque is written as: NG ^r I^r; = 3GE r3 (14) where ^r = hri is the unit length vector indicating the geocentric direction to the satellite’s centre of mass, and r = jrj is the geocentric distance of the satellite. GE is the Earth’s gravitational constant. The remaining environmental torques were derived from the following equations: 0 (! I !) =I 1 0 N 0 N =I 1 G d ! @ N =I A = + @ (! I !) =I A , @ NG dt (! I !) =I NG N =I x xx x xx y yy y yy z zz z zz =Ixx ;y =Iyy ;z =Izz ;x 1 A; (15) where all terms on the right-hand side have now been determined. Figure 37 illustrates these corrections for one orbit. 8.2.4. The satellite dynamical model, III: The solar radiation torques The solar radiation reflected and absorbed by the outer surface of the satellite caused a force working on the satellite which was not balanced with respect to its centre of mass, and therefore resulted in torques. The size of these torques depended on the orientation of the satellite with respect to the direction to the Sun, which was determined by the nearly fixed angle between the satellite’s z axis and the Sun ( = 43 ), and the scan-phase , which described the angle between the x axis of the satellite and the great circle through the direction to the Sun and the satellite’s z axis (see Figure 12). Because the satellite had a basic three-fold symmetry (see Figure 3), the torques caused by the solar radiation were predominantly a function of 3 . The periodicity at multiples of 3 applied when the torques were seen in a y z ] be such a reference frame fixed with respect to the solar direction. Let [ x on the great circle reference frame with z along the satellite nominal spin axis, x = z x through z and the direction to the Sun as seen from the satellite, and y 263 THE HIPPARCOS MISSION Table VIII The solar radiation related torque components as observed in the accelerations. The observed values, after having been corrected for Sun–Earth distance variations, were fitted to a linear function of time with day 800 as reference point. The coefficients (at day 800) are measured in mas s,2 and the derivatives in mas s,2 year,1 . Axis Term x sin sin 2 sin 4 sin 5 sin 7 x x x x Coefficient ,2 28005 0 00031 0 59429 0 00012 ,0 20310 0 00008 0 09826 0 00008 ,0 01112 0 00009 ,2 16908 0 00024 ,0 64572 0 00011 ,0 17513 0 00008 ,0 09407 0 00010 0 00948 0 00011 ,0 77013 0 00022 ,0 12250 0 00008 ,0 02593 0 00006 ,0 01035 0 00006 cos cos 2 cos 4 cos 5 cos 7 y y y y y sin 3 sin 6 sin 9 sin 12 z z z z : : : : : : : : : : : : : : : : : : : : : : : : : : : : Derivative ,0 00980 ,0 00895 ,0 00028 ,0 00230 : : : : 0:00039 0:01662 ,0 00968 : 0:00319 0:00099 0:00152 , 0:00261 0:00337 0:00049 0:00106 0 00042 0 00016 0 00010 0 00011 0 00012 0 00032 0 00015 0 00011 0 00013 0 00014 0 00029 0 00011 0 00008 0 00007 : : : : : : : : : : : : : : completing the triad. The solar radiation torques were approximated in this frame by: 0 0 1 m 0 a sin 3n 1 n N R @ b0 A + X @ bn cos 3n A n=1 cn sin 3n 0 (16) N The torques R as experienced in the satellite reference frame were obtained through a rotation by around the z axis: 0 cos NR = @ , sin sin cos 0 0 1 0 0A 1 N R As a result, the solar radiation torques around the satellite approximated by: NRx b0 sin + 12 Pmn=1(an , bn) sin(3n , 1) + 1 Pm (a + b ) sin(3n + 1) 2 n=1 n n (17) x and y axes were (18) and: NRy b0 cos , 12 Pmn=1(an , bn) cos(3n , 1) + 12 Pmn=1(an + bn ) cos(3n + 1) (19) 264 F. VAN LEEUWEN Figure 38. The primary components of the solar radiation torques as observed in the accelerations around the three main satellite axes. The modulation has a period of a year and results from the varying distance between the Earth and the Sun, due to the ellipticity of the Earth’s orbit. This defined the coefficients characteristic for torques caused by the solar radiation. A further peculiarity of these torques was the fact that their strength varied inversely proportional to the square of the distance between the Earth and the Sun, i.e. by about 3:5 per cent. This can be seen in Figures 38 and 39, which show the main torque components (expressed in the measured accelerations) as observed over the mission before and after correction for the varying distance to the Sun. THE HIPPARCOS MISSION 265 Figure 39. The primary components of the solar radiation torques, corrected for the varying distance between the Earth and the Sun. The solar radiation torques around the x and y axes were of the order of 2 Nm, which translated into accelerations around these satellite axes at a level of 2.3 mas s,2 . Around the z axis the torque (or acceleration) was about a factor 3 smaller. Table VIII shows the amplitudes of the main solar radiation related torque components. The observed changes in the solar radiation torque coefficients probably relate to changes in the light reflection and absorption by the satellite and its solar panels due to the exposure to the space environment. 266 F. VAN LEEUWEN The coefficients in Table VIII can be transformed back into the coefficients an , bn and cn as given in Equation 16, using the inertia tensor values given in Table VII. From a comparison of the results obtained for the x and y axis for the an and bn coefficients it is clear that although the results are in general agreement, additional contributions not described by Equation 16 also play a role. This was also noted from the presence of various other harmonics, most noticeably a 1st harmonic ( ) around the z axis, for which the amplitude increased much towards perigee. These coefficients revealed the magnetic moment of the satellite and its interaction with the Earth’s magnetic field. 8.2.5. The satellite dynamical model, IV: The magnetic torques The torque caused by the interaction of the magnetic moment vector Earth’s geocentric magnetic flux density vector is given by: B N M = m B m and the (20) As seen from the satellite reference frame, is rotating. In the satellite reference frame the flux density vector is expressed as: 0 cos B = @ , sin 0 B 1 m sin m 0 m cos m 0 A B 0 (21) 1 B where m describes the instantaneous orientation of with respect to the satellite x axis. Within the satellite reference frame was assumed constant. This gave the following relations: m 0m B ,m B 1 y z z y NM = @ mz Bx , mxBz A mxBy , my Bx (22) where at apogee the torques around all three axes were approximately modulated by . Away from apogee this modulation became disturbed due to the changes in , and both the amplitude and the period were varying strongly. The description for used the 1985 coefficients of the International Geomagnetic Reference Field (IGRF) (Barraclough, 1985), describing the magnetic field as a series of spherical harmonics. The most important (dipole) coefficient is proportional to r ,3 . The magnetic torques were smaller on average than the gravity gradient torques, but increased in a similar fashion with decreasing distance to the Earth. Two magnetic moment components were observed: along the y and z axes of the satellite. Their behaviour over the mission is illustrated in Figure 40. The magnetic moment along the z axis changed significantly during eclipse periods, when the power supply switched from solar panels to batteries. Studies of the wiring of the spacecraft could not reveal the likely causes for these effects. The magnetic torques would have been insignificant for the nominal mission. B B THE HIPPARCOS MISSION 267 Figure 40. Measurements of the magnetic moments of the satellite. The y component was most clearly present. There was probably also a significant z component during normal operations (middle graph) and there appeared to be a significant z component during eclipse conditions (bottom graph), which could, however, only be solved for under good attitude convergence conditions, and had to be fixed otherwise. No corrections were applied for compression or stretching of the Earth’s magnetic field due to interaction with the solar wind. The analysis of the data indicated, however, that such effects were probably present. The analysis results for torque components that are likely to be related to the magnetic moment of the satellite show a modulation with the precession phase of the satellite’s z axis as well as 268 F. VAN LEEUWEN Figure 41. Three coefficients, showing through their variability a possible link with the magnetic moment of the satellite and a distortion of the Earth’s magnetic field at high altitudes. The short period variation is probably related to the spin period of the satellite’s z axis (57 days). The long period modulations could be related to the time interval between two conjunctions of the Sun and the satellite at apogee (588.7 days). with the orientation of the satellite’s orbit with respect to the direction of the Sun as seen from the Earth, that may well reflect these compressions (Figure 41). 8.2.6. The satellite dynamical model, V: Miscellaneous effects Among the various miscellaneous effects noted in the torque analysis was a local warming-up during perigee passages, followed by local radiation and a resulting THE HIPPARCOS MISSION 269 Figure 42. Short term variations for the constant for the y axis torques at a time of very low perigee passages. The orbits with even numbers are connected to show the effect of earth longitude on this coefficient. The period of the modulation is 8 days, equivalent to 18 orbits. windmill effect. This effect was very sensitive to the density of the outer layers of the atmosphere at perigee, and occurred only during very low perigee passages (400-500 km altitude). The warming-up of the spacecraft during these passages was clearly noticed from the drift of the on-board time (OBT, see Section 3.1). Considerable variations appeared to exist in the atmospheric densities as a function of Earth longitude: clear correlations were observed for the constant coefficient in the y axis torques between perigee passages at nearly the same longitude. As can be seen from Figure 1, perigee longitudes for alternating orbits were much closer related than for successive orbits. Figure 42 shows the evolution of the constant coefficient for the y axis solutions over a period of low perigee height, illustrating this effect. Some gyro de-storage exercises produced a considerable additional torque on the satellite, depending on the orientation of the angular momentum vector of the gyro with respect to the spin axis of the satellite. This allowed a calibration of the angular momentum for gyros 4 and 5 to be made. The observed offsets showed that orientation parameters were as expected, while the gyro angular momentum appeared to be 4 percent higher than specified. Finally, impulse like torques were experienced occasionally as result of accidental hits of the satellite by micrometeorites. Most hits were very minor, but two reached a level comparable to a thruster firing, producing a change in rotation rate of a few arcsec s,1 . Such occasions were recognized from the routine inspections of the gyro data by NDAC as a discontinuity in the measured rotation rates that was not associated with a thruster firing. 8.2.7. The attitude reconstruction, NDAC approach The attitude reconstruction as performed by NDAC was based on the torque calibrations and modelling described in the preceding sections. The descriptions of the torques were used to model the accelerations from which, through a numerical integration of the Euler equations, estimates for the inertial rates were obtained. 270 F. VAN LEEUWEN These integrations took place over intervals between thruster firings. Integrations across thruster firings used the calibrated thruster responses (see Figure 27) and provided starting points for the inertial rates for the integration over the next thruster firing interval. The zero-point offsets of the inertial rates were calibrated using the gyro data. The corrected rates were integrated to provide a first approximation of the attitude angles. These approximate angles (given at three points for every observational frame) were used to predict star mapper transit times. The differences with the observed transit times were expressed as low order polynomials, which, together with the model that served as a first approximation, provided the actual attitude reconstruction. The main vulnerability of this method appeared to be around thruster firings where sometimes, due to a lack of star mapper observations, insufficient constraints for the solutions were available. A detailed description of the NDAC attitude reconstruction can be found in Chapter 7 of Volume 3 of (ESA, 1997). 8.2.8. The attitude reconstruction, FAST approach. I: the model The FAST approach was based on a simplified integration of the Euler equations while the satellite was near apogee, and a polynomial approach closer to perigee. The simplified model shows the harmonics of the solar radiation torques to appear in the modelling of the attitude angles, accompanied by discrete functions over thruster firing intervals, which described the effects of the thruster firings on the attitude. Pre-launch research aimed at reducing the modelling error under normal geostationary conditions, while using a minimum number of degrees of freedom. This model worked well in apogee conditions, but was no longer adequate near perigee. The use of the rigid body equations to completely describe the satellite motion was excluded by FAST as it was considered too difficult to calibrate with the required accuracy the telescope reference frame position with respect to the inertial axes of the satellite and to maintain such an accuracy during all the satellite life. 8.2.9. The attitude reconstruction, FAST approach. II: the attitude angles The FAST attitude representation was described relative to the reference great-circle frame (Section 8.5), defined by the position of its pole in ecliptic coordinates, R and R . This pole was selected as the position of the z axis in the nominal scanning law for a reference time halfway between the start and the end of the data-set concerned. Relative to this reference system, three attitude angles were defined: , and . Particular attention was devoted to the choice of the attitude angles. In fact, the various sets of three angles are related among themselves by non-linear transformations so that their choice influences the simplicity of the attitude model, the estimation method, and accuracy. The chosen angles are related to the classical 3-2-1 Euler angles e , e and e by: = e ; = e ; = sin sin e , arctan cos + cos (23) THE HIPPARCOS MISSION 271 The chosen representation of the angle was motivated by the fact that, as its rate of change is much closer to the inertial rotation rate around the satellite z axis, it improved the modelling of the scanning motion. Experiments showed that estimation results obtained with these angles were a significant improvement upon those obtained by using e . 8.2.10. The attitude reconstruction, FAST approach, III: Approximate motion equations The , , attitude angles were modelled as the outputs of the following two independent systems of differential equations, which were obtained as an approximation of the satellite dynamic equations in the case of small perturbations. ! _ z = z (t) _ = ! z (24) and: ! _ x = x (t) !_ y = y (t) _ = _ = ,!0 + ! x !0 + !y (25) where !0 is a constant parameter assumed known and equal to the satellite’s nominal spinning rate. The functions x (t), y (t), z (t) are the derivatives of the rate components !x , !y , !z . In a first rough approximation they model the effect of the torques acting on the satellite, but more precisely they have been introduced not only to model the torque effect but also to compensate the approximation introduced by using the above simplified equations to model the telescope motion. Considering that any given set of three continuous time functions (t), (t), (t) can be obtained as the output of the above dynamic system by the application of a suitable set of input functions x (t), y (t), z (t) and by initial conditions, the model error could be reduced as much as required by attributing a sufficient number of degrees of freedom to the input functions. 8.2.11. The attitude reconstruction, FAST approach, IV: System Input Functions The system inputs x;y;z (t) were decomposed into the sum of two terms, the control inputs ux;y;z (t) and the disturbances dx;y;z (t) respectively. The control inputs were introduced to describe the effects of thruster firings on the on-board attitude control: they were assumed to be a sequence of ideal pulses (Dirac functions) of unknown amplitude applied at the central time of the thruster firings. The disturbances were introduced to describe the effects produced by the perturbing torques and to keep into account, at the same time, the difference between the above simplified dynamic system and the reality. The disturbances were expressed by different 272 F. VAN LEEUWEN Figure 43. Flow charts for the FAST attitude reconstruction models. mathematical models (Fourier and polynomial models) depending on the length of the time interval considered and the satellite perturbing conditions. In the case of Fourier models the disturbances were described by a Fourier series with unknown coefficients and the response of the system was computed on the basis of the above differential equations. In the case of polynomial models, the system response to the disturbances was directly described by polynomial series. The resulting attitude model is presented for the two cases in Figure 43. The following different models have been used: long-term Fourier model: this was the standard model, used in normal operating conditions over a time interval corresponding to a full rotation around the satellite z axis, 128 min. Independently from the actual duration of the considered time interval, the disturbances were described by a Fourier series whose fundamental harmonics had the frequency !0 , corresponding to the satellite’s nominal spinning rate used in the approximate equation of motion (Equation 25). In this particular case the forced solution of the considered differential equations had the following expression: d (t) = bs0 t sin !0 t + bc0 t cos !0 t + Pn k=2 bsk sin(k!0 t) + bck cos(k!0 t) d (t) = cs0 t sin !0 t + cc0 t cos !0 t + Pn 2 csk sin(k!0 t) + cck cos(k!0 t) k = Pm (26) d (t) = k=1 ask sin(k!0 t) + ack cos(k!0 t) THE HIPPARCOS MISSION 273 A variable number of harmonics, in the range 12 to 36, was used in modelling , (m), and in modelling and , (n). The number of harmonics was determined by an adaptive method based on the computation of the FisherSnedecor test (Cramér, 1946). The adaptation was made by increasing the number of harmonics by 3 at a time, starting from 12. The optimal number was the maximum number for which the Fisher-Snedecor test value was still above a given threshold; intermediate term Fourier model: this model was used over intervals of 30 min to 115 min. The perturbing torques were modelled like above by a finite Fourier series of unknown amplitude harmonics (Equation 26), but without the mixed terms for and (describing the torque caused on through the change of and vice versa). Here the fundamental period was chosen equal to the length of the time interval, and the minimum number of harmonics at the start of the computation was determined by the interval length. They were increased by one at a time, followed by the tests described above, until a test value within a given threshold was obtained; short-term polynomial model: this model was used for the shortest considered time intervals, viz. a period between two thruster firings (varying from a few minutes to more than half an hour, but on average about 12 to 15 minutes). The perturbing torque effect was modelled by Legendre polynomials, the degree of which was adapted to the considered case, independently for and for and . The polynomial degrees were increased by one at a time until the Fisher-Snedecor test value obtained was within a given threshold; short-term polynomial model in penumbra: while the satellite was in penumbra, before and after an eclipse, the solar radiation torque followed a transient from high values to very low or vice versa. Conditions during different transients were similar but never the same. The transient was modelled as a stochastic process for which a Karhunen-Loève expansion (Papoulis, 1991) was computed. This provided an ordered set of orthonormal functions, of which each one was approximated by a Legendre polynomial representation. The first 10 ( ) and 15 ( and ) functions were stored. The order of the Karhunen-Loève expansion used was adapted to the particular case considered, starting from a minimum value determined by the number of star mapper transits available, up to the maximum number of stored functions, applying at each step the Fisher-Snedecor test. The numerical methods used for the FAST attitude reconstruction are detailed in Chapter 7 of Volume 3 of ESA (1997). 8.2.12. Accuracy assessment The accuracy assessment was primarily based on comparisons between FAST and NDAC reduction results, as carried out at the Centre for System Studies (CSS) in Torino by F. Donati (Donati & Sechi, 1992). The comparisons also involved 274 F. VAN LEEUWEN Table IX Estimated standard deviations of the random terms in the differential rotation in body axes, expressed in milliarcsec. Column (a): all frames included; column (b) excluding outliers. Component Estimates (mas) (a) (b) 129 81 12 90 47 10 82 54 24 14 3 3 improvements in the along-scan attitude as obtained with the great-circle reductions (see Section 8.5). Before comparisons could be made, the data obtained from NDAC and FAST had to be synchronized and rotated to coincide. The rotation matrix reflected differences in the Input Catalogue positions (due to independent updating mechanisms, see Section 8.3) and differences in the star mapper geometry calibrations. The remaining standard deviations between the results of the two consortia were examined for three coordinates: 0 1 = @ A ; (27) where (, , ) describe the rotations around the x, y and z axes respectively. Denoting by 1q and 2q the standard deviations of the two attitude reconstructions to be compared, in any of the angles (q = , , ), the following holds: q2 = 12q + 22q , 21q 2q ; (28) where is the correlation coefficient between the two realizations of the attitude. The correlation was close to one, as both solutions used the same star-mapper transit data. Table IX shows the results of the last attitude comparison, where it should be noted that the data for the coordinate had been improved by the great-circle reductions. As is very clear from these results and the discussion in Section 8.2.1, the projection errors in the great-circle reduction that resulted from the errors on the reconstruction of the z axis position (angles and ), were for most of the time negligible. There were, however, occasions when a very small amount of data had much larger uncertainties in the attitude determinations. These occurred very close to thruster firings, where control over the attitude reconstruction was partly lost, and at times of very high background signals in the star mapper detectors, when the numbers of available transits were much reduced. Most of these problems were detected and taken care of by the great-circle reduction process. THE HIPPARCOS MISSION 275 8.3. EARLY IMPROVEMENTS OF THE INPUT CATALOGUE The processes described in this section made use of the star mapper and attitudereconstruction data, and had only indirect influences on the mission results. The techniques used, however, may have much wider applications. The aim of the improvements to the catalogue was to obtain at an early stage improved star positions for the epoch of observation, leading to better results for the on-board as well as the on-ground attitude reconstruction (for which the accuracies as provided in the Input Catalogue were inadequate), to obtain improved photometric data required for various calibrations, to detect misidentifications early in the mission, and to provide much improved input positions of double stars to the main-mission double star processing (where absolute positions were often very poorly known). The processes as described here were implemented by NDAC. Positional improvements were also implemented by FAST, in particular as obtained from provisional sphere solutions and as part of the final iteration between the attitude reconstruction and the great-circle reduction, sphere solution and astrometric parameter determinations. The accuracy of the attitude reconstruction and the star mapper transit times was about 5 to 10 times higher than the average accuracy of the positional information in the Input Catalogue. As a result, the residuals obtained during the first runs of the attitude reconstruction processes reflected mainly the errors in the star positions and not the noise on the transit time determinations. By collecting sufficient numbers of these residuals, positions could be adjusted. Accurate photometric data were available for only a small fraction of the Input Catalogue stars. The intensities obtained from the star mapper processing were used to improve this situation as early in the mission as possible, and to provide accurate colour information for a wide range of calibration processes. 8.3.1. The updating mechanism The updating mechanism used a running least squares solution, based on Householder transformations as described by Bierman (1977). The advantage of this method is the way the normal equations are represented: as an upper triangular matrix with dimensionally the same form as the observation equations. Thus, an old solution can be combined with new observations simply in the form of a set of additional observation equations. The positional and photometric updates for each star were given by these normal equations stored in a working catalogue. For example, the normal equations for the positional updates described the differential corrections to the current catalogue positions. The updates for the two positional coordinates were linked, as all information came from one-dimensional observations, giving correlated corrections to right ascension and declination. The provision of vertical and inclined slits effectively produced a two-dimensional position for a transit through both slit groups, which was the way positional updates were implemented by FAST. For the magnitudes the information consisted of sums of weights and weighted residuals. The accumulation of the information on the 276 F. VAN LEEUWEN magnitudes was done in a pseudo intensity scale to avoid a bias in the accumulated results. In order to avoid faulty data to enter these updates, a minimum of 8 to 10 observations was required before a first solution was made. Before this minimum number of observations was reached, observations were put into a storage file, from where they could be retrieved when more data for the same star became available. Data for single stars was processed without supervision, while data for double stars (for which component confusion was not uncommon) were processed in an interactive mode first. 8.3.2. The use of the early improvements The use of the positional improvements was described by van Leeuwen et al. (1992a) and Turon et al. (1992b). Positional improvements for over 15 000 stars were already obtained before the start of the routine data reductions, using the 1300 hours of data sent from ESOC to the consortia for testing interface and reduction software (Lindegren et al., 1992a). Much improved positions for some 60 000 stars had been obtained before the first sphere reconstruction results became available. A small number of these involved corrections of over 2 arcsec, at which level the pointing of the instantaneous field of view for the main-mission data became affected. These updates were incorporated in the star catalogue used for the mission operations, and led to a significant improvement in the on-board attitude control, with much less influence from position errors on the attitude results. For a small number of stars the positional corrections were so large that only observations obtained after the corrections to the Input Catalogue had been applied could be used in the preparation of the final catalogue results. The star mapper processing for double stars by NDAC resulted in very much improved positions and in some cases improved separations and orientations for 1643 systems unresolved in the Input Catalogue and with a minimum separation of 1.5 arcsec. For all unresolved double stars with separations above 1.5 arcsec the components were entered in the working catalogue: the brighter component under the original HIC (Hipparcos Input Catalogue) number of the system, and the fainter component with a new number. The initial positions were set by the separation and orientation parameters and the original position for the HIC number. The two components were treated as two separate stars in the star mapper processing, but with the knowledge of the proximity of the other component. The star mapper double star results were of considerable help in the processing of the main-mission double star information, which, due to slit ambiguities, had often problems with setting proper starting parameters for solutions (Söderhjelm et al., 1992). Figure 44 compares separations and orientations as given in the Input Catalogue with the data derived from the star mapper processing. The catalogue updating process accumulated photometric data for over 90 000 stars, which were used to improve the colour index information required in several calibration processes. Figure 45 shows the standard errors of the individual photo- THE HIPPARCOS MISSION 277 Figure 44. Star mapper results for double stars, compared with the Input Catalogue values. A total of 1643 stars that were entered as a single component in the Input Catalogue were resolved using star-mapper data. 278 F. VAN LEEUWEN Figure 45. The standard errors for the accumulated BT and VT magnitudes for constant stars as a function of magnitude. The solid diagonal lines show the expected relation for Poisson noise. The increase in noise towards fainter magnitudes is due to background contributions, towards brighter magnitudes to modelling inaccuracies and saturation. metric star mapper observations. Errors on the mean values for constant stars were about 10 times smaller. The photometric processing of the star-mapper data was very similar to the Tycho photometric reductions, described in Section 8.8.3. 279 THE HIPPARCOS MISSION 8.4. IMAGE DISSECTOR TUBE DATA REDUCTION The task of the image-dissector-tube-data reductions was to determine the modulation parameters for the transits of program stars passing over the main grid. Observations were collected over time spans of an observational frame, 32/15 s, containing 2560 sampling periods of 1/1200 s. The modulation parameters described the mean intensity of the signal and the amplitudes and phases of the first and second harmonics of the modulation. In this section the methods used to derive these modulation parameters are described, as well as statistical tests and calibrations performed on the results. The reduced data formed the input for the great-circle reductions, the double star analysis and the main-mission photometric reductions. 8.4.1. The signal model The estimation procedures used a reference phase pk for each sample k . This reference phase measured the local modulation phase relative to a phase equal to zero at mid-frame time at the fiducial reference line on the main grid, halfway between slits 1344 and 1345 (see Figure 5). Methods used to assign precise reference phases pk to each sampling have been described in detail in Chapter 5 of Volume 3 of ESA (1997). Using the attitude reconstruction results, the local scan velocities for each field of view were calculated, and together with the projection-related distortions of the grid (which was accounted for by describing the stellar positions on the grid in direction cosines) and the catalogue position of the star (transformed to the epoch of observation and corrected for aberration), the relative phase for each sampling was calculated. The grid positions obtained this way were compared with the grid positions for the positioning of the instantaneous field of view (IFOV), as derived from the coil current values for the observation and the coil current calibration matrix. If the discrepancy was more than 10 arcsec, the observation was rejected. A record was kept of the observed differences as an overall quality control parameter (see Figure 23). With the precise relative phase estimates for each sample available, the processing could start. The signal model described the sequence of statistically independent and Poisson-distributed counts Nk as having ideally a time-periodic behaviour with an expected value of E (Nk ) Ik . Various representations of the modulated signal have been used, but only one of these was implemented in the actual fitting of the data: b) = b1 + b2 cos pk + b3 sin pk + b4 cos 2pk + b5 sin 2pk : (29) All other representations were derived from this “b-parameters” fit through transIk ( formations. The expression in parameters with direct physical interpretation was given by: Ik = Ib + Is 1 + M1 cos(pk + g1 ) + M2 cos 2(pk + g1 + g2 ) ; (30) where M1 and M2 are the modulation amplitudes and g1 and g1 + g2 the modulation phases for the first and second harmonic. The signal and background intensities 280 F. VAN LEEUWEN were given by Is and Ib respectively. The model as used by FAST was closely related to this representation: I (a) = a1 + a2 cos(p + a3 ) + a4 cos 2(p + a5 ): k k (31) k The reference phase used by FAST was based on a weighted mean of the first and second harmonic phases: g0 = 0:75a3 + 0:25a5: (32) For single stars the amplitude ratio of the two harmonics was given by = a4 =a2 and their phase difference by = a5 , a3 , where (, ) are instrument parameters that are functions of star colour, field of view and position on the grid, as well as time. In NDAC a model was used which represented the second harmonic relative to the first harmonic parameters: I ( )= k 1 + 2 cos(pk + 3) + 4 cos 2(pk + 3) + 5 sin 2(pk + 3) : (33) In this representation, 4 and 5 are the instrument parameters that describe the amplitude ratio and phase difference of the first and second harmonic in the modulation (see Section 8.4.5: The OTF calibration). In the NDAC reductions, g0 = 3 was used as reference phase. Next to these five-parameter models, three parameter models were used. These three-parameter models were obtained through the implementation of the instrument parameter calibration values. In the NDAC representation the three parameter model was simply given by ( 1 , 2 , 3 ), as the remaining two parameters were not used in defining the signal phase and amplitude. In the FAST representation the three parameter model was given by: I (r) = r1 + r2 cos(p + r3 ) + r2 cos 2(p + r3 + ): k k k (34) The three parameter models were important in the recognition of double stars. In a signal composed of the contributions of more than one star, the calibration values for (, ) or ( 4 , 5 ) no longer applied (see Section 8.4.6). 8.4.2. Relations for double and multiple images The most important parameter relations were those between the coefficients in Equations 29 and 30, as they allowed the interpretation of signals obtained from single and multiple images. Assuming average modulation coefficients for all 1 and M 2 , and the same reference phases pk components in a multiple image as M 2 for all components, the following relations were derived (where the parameters M and g2 were eliminated using the OTF calibration): b1 = I + X I s;j b j 1 b2 = M X 281 THE HIPPARCOS MISSION Is;j cos g1;j j 1 b3 = M X Is;j sin g1;j j 1 b4 = M X Is;j cos 2(g1;j + ) j 1 b 5 = M X Is;j sin 2(g1;j + ); (35) j This was used (in a somewhat different form) as a test for single stars (see Section 8.4.6) and as the basic equations for the double star investigations (see Section 9.1). 8.4.3. Data binning At the time of the development and early implementation of the data reduction software, it was essential to look for ways of limiting the time spent on the processing of the 1200 IDT samples received every second. Both data reduction consortia opted for a binning strategy, reducing the main processing tasks by a very considerable factor in both computing time and complexity. The binning strategy assigns a reference phase pi to all samples with phases pk falling in a phase interval pi , dp to pi + dp. With pk = pi + pkji , and ni samples falling in bin i, the expected count in bin i can be expressed as a modification of Equation 29: 1 X ni Nkji = b1 + b2 n1 i cos pi +b3 n1i sin pi P P P +b4 n1i cos 2pi +b5 n1i sin 2pi P cos pkji , sin pi cos pkji + cos pi P P cos 2pkji , sin 2pi cos 2pkji + cos 2pi sin pkji sin pkji P P sin 2pkji sin 2pkji (36) The advantage of the binning resulted from the pk being small and the values of pi being fixed, thus avoiding repeated trigonometric calculations. Two different implementations were used: the FAST consortium applied a strategy using 64 bins, and neglected the effect of pk , in which case binning corresponded to a quantization of the relative phase. It was estimated that the error on the relative phase pk resulting from the binning should be less than 10 mas, in which case the quantization effects on the grid phase estimate would be negligible with respect to the photon noise. The quantization noise q for l bins was given by: q = ps 2 3l ; (37) 282 F. VAN LEEUWEN where s is the grid-period of 1.2074 arcsec. For l > 35 it follows that q < 10 mas. l was rounded up to the nearest power of 2, i.e. l = 64, giving q = 5:5 mas and a noise contribution to the grid phase estimates less than 0.5 mas. NDAC performed a substantial number of simulations and tests with various binning strategies, using both noise-free and Poisson-noise simulated data. The results of these tests showed that a substantial reduction in computing time without any significant loss of accuracy could be obtained using l = 12 bins and carrying along up to second order corrections for pk . The noise contribution to the phase estimates resulting from the binning this way, was less than 0.2 mas. In the FAST approximation Equation P 36 reduces to Equation 29, with pk being replaced by the nearest pi and Ni = Nkji being the total count per bin. In the NDAC approximation the equations obtained were: 1 X ni Nkji = b1 + b2 n1 cos pi i P (1 i k i P pkji , 12 p2 j ) + cos p P p j P P 1 +b4 cos 2p (1 , 2p2 j ) , sin 2p 2 p j P P 1 sin 2p (1 , 2p2 j ) + cos 2p 2 p j : +b5 1 +b3 n i sin pi i ni ni i P , 12 p2 j ) , sin p (1 k i k i k i i k i i k i i k i (38) The relevant trigonometric terms were calculated once only and stored in memory. For certain values of the scan-velocity the relative grid phases pk would assume only a limited number of different values. At nominal scan velocity (168.75 arcsec s,1 ), the phase shift from one sample to the next was 41.92 . At a scanning speed of 170.456 arcsec s,1 , the phase shift to the next sample was 42.35 , which meant that the phase shifts of 17 samplings fitted in exactly 2 complete modulation cycles. Similar situations happened for scan velocities of 173.866 arcsec s,1 (25 samplings in 3 cycles) and 167.178 arcsec s,1 (26 samplings in 3 cycles). In the case of 64 bins without phase correction, this left a large number of bins unoccupied, and could cause systematic differences between assumed bin-phases and actual mean bin phases. The latter effect was, however, small with respect to other sources of noise, most notably the Poisson noise on the photon counts. The scan velocity was changing continuously, and a resonance situation like that described above would normally not persist over more than a few frames. No measures were taken to remedy these phenomena by FAST, while in NDAC they were implicitly taken care off in the standard reductions. In FAST, the statistical index T (Section 8.4.4) was able to detect non-negligible biases in the grid phase such as due to resonances. A full description of the solution techniques used for the binned equations can be found in Chapter 5 of Volume 3 of ESA (1997). Solutions were based on weighted least squares applied iteratively, using in the final solution the previous intensity estimates as an indicator for the Poisson noise on the photon count. THE HIPPARCOS MISSION 283 8.4.4. Statistical tests Two test parameters were derived at this stage: F and T . The first of these, F , tested the null hypothesis that no modulation was present in the signal, in other words, that the observed photon counts collected for a transit were a stationary white Poisson noise; hence F appeared significant when the photon-counts were sufficiently modulated. The second of these, T , tested the null hypothesis that the residuals left after application of the five-parameter model represented a zeromean stationary white noise. Hence, T appeared significant when the residuals were modulated (failure of the 5-parameter model) or the original photon counts were not Poisson distributed. The tests were carried out on the binned counts. In NDAC, a 2 value for the T statistic was derived for the binned mean counts, ^ ): assuming them to be represented by Ik (b i2 h1 P ^) N , I ( b k k i n 2 = 1 P Nk i n i=1 l X j i i (39) j where l = 12 is the number of bins and where the denominator represents the expected variance of the residuals. The 2 value was transformed into a pseudoGaussian variable, which under the null hypothesis should have a unit-normal distribution: 3 s 2 2 !1=3 T = 92d 4 d , 1 + 92d 5 (40) where d is the degrees of freedom, normally equal to l , 5 = 7 for the T statistics. Transits for which jT j > 4:0 were flagged in the Epoch Photometry Annex file (hip ep.dat) as suspicious. In FAST, a binned mean square error was computed as: h i2 ^) l N , I (b X i i s2 = (41) ~ i=1 (l=m)Ii (b) where l = 64 is the number of bins and m the total number of samples used. The statistic T was then computed as: s " # l , 5 s2 T = 2 l,5 ,1 (42) Under the null hypothesis, s2 follows the 2 distribution with l , 5 = 59 degrees of freedom and T approaches asymptotically the unit-normal distribution. The null hypothesis was rejected if jT j > 3:5, which happened in less than 0.5 per cent of all transits. 284 F. VAN LEEUWEN In NDAC, a 2 value for the F statistic was derived for the binned mean counts, assuming them to be stationary: 2 = h i X P NP , b N 1 l 1 =1 1 2 (43) k ji ni i ^ k ji ni As for the T statistic, the 2 value was transformed into a pseudo-Gaussian variable, except that the degrees of freedom was d = 11. In FAST the binned mean square of the estimated modulation components was computed as: p2 = X i h i I (b^ ) , ^b1 =1 (l=m)I (b~ ) l i 2 (44) i having four degrees of freedom. The statistical index F was then defined as a Fisher ratio between the statistics in Equations 44 and 41: 2 F = s2=p(l=,4 5) (45) Since (l , 5) was rather large, 4F almost followed a 2 distribution with four degrees of freedom. Accordingly, the null hypothesis was accepted (no modulation found) if F < 4:5. This happened in less than 1.5 per cent of all transits. 8.4.5. Calibration of the Optical Transfer Function The Optical Transfer Function (OTF) described the characteristics of the modulated signal over both fields of view. The calibration of the optical transfer function used the values for (; ) or ( 4 , 5 ) collected over one orbital period of the satellite. The calibrated values were used in the 3-parameter model described in Section 8.4.6. The modulation parameters were associated with a position on the grid (G; H ) at mid-frame time, and with the colour index for the star observed. The instrument parameters ( 4 , 5 ), obtained for stars not known or found to be double, were modelled with a two-dimensional third order polynomial in position, first and second order in colour index, and cross-terms between colour and position on the grid, 14 parameters in total. Every calibration covered the data collected during one orbit. The calibrations were done independently for the two fields of view. Of the parameters, the positional dependence was the most important. An example of this dependence is shown in Figure 46. The calibration was time dependent, as is shown in Figure 47, and affected by refocussing of the instrument. The colour dependence in the preceding field of view for 4 increased during the mission by a factor 2, while for 4 in the following field of view and for 5 this dependence was much less a function of time. The amplitude ratio 4 decreased towards redder stars, but could not be calibrated for the very red stars, as too few measurements THE HIPPARCOS MISSION 285 Figure 46. The dependence of 4 and 5 on the grid position at the beginning of the mission. The scan direction is indicated by the arrow, the two graphs for 4 and 5 refer to the two fields of view. Scales for the two fields of view are identical. were available. At V , I = 2 the decrease was approximately 10 per cent in the following field of view, less in the preceding field of view. For extreme-red stars these calibrations were not possible, while an extrapolation of the calibrations was very unreliable. Duplicity indicators for such stars should therefore be treated with caution. In FAST, the calibration of the optical transfer function was, as were most calibrations, performed twice. An OTF calibration was done in Utrecht by the first look task once a week. The mean for a calibration period was made and used by the IDT data processing as described in this chapter. In a second run, the five 286 F. VAN LEEUWEN Figure 47. Evolution with time for the phase difference and the amplitude ratio . Data for the preceding field of view are shown by filled symbols (points in the lower parts of the graphs), for the following field of view by open symbols (points in the upper parts of the graphs). Features in the top graph around days 450 and 1020 are related to variations in the satellite’s exposure factor, shown in Figure 2. parameter solutions obtained in this processing were used in a more refined analysis intended to be used by the multiple star and photometry tasks. About twenty orbits contributed to each calibration period, chosen among those giving the best results in the great-circle reduction. The procedure was as follows. All the results of the five parameter solution were examined and several tests performed in order to exclude known double stars, faint 287 THE HIPPARCOS MISSION stars (magnitude > 11), stars with an unknown colour, and all stars for which either the ratio of the ‘ac’ to ‘dc’ photometry or the difference a3 , a5 suggested that the star was double. Additional rejections were made from the analysis of histograms of these quantities. The ‘ac’ intensity was computed and used to determine M1 and M2 from a1 , a2 and a4 . Each quantity M1 ; M2 and a3 , a5 for each field of view was expressed as a third order polynomial in the grid coordinates G and H plus a similar polynomial multiplied by C , 0:5, where C is the colour index; this gave 20 parameters in total. The equations were solved by least squares giving for each a weight proportional to the intensity of the star (a1 ). In addition, the reference intensity response was calibrated for a mesh of 19 19 points on the grid for three classes of colours and also by a polynomial of the third order in G and H , second order in colour and some mixed terms. This formed part of the FAST photometric calibrations described in Section 9.3. 8.4.6. The Three Parameter Solution In NDAC, the optical transfer function was applied to the (G; H ) coordinates and star colour of each frame transit to provide predicted values for 4 and 5 . These provided, together with the relevant elements of the covariance matrix ,1 , a 2 estimate for the likelihood of the signal being the result of one point-source only: B 2 = B,1 ]44 + 2 2 4[ B,1 ]45 + 52 [B,1 ]55 4 5[ (46) where 4 is the difference between the observed and predicted value of 4 , and similarly for 5 . The 2 values were collected per star for the purpose of double star recognition. This was done using a transformation of 2 into a variable that has a flat distribution for single point-source objects, but becomes increasingly skew for disturbed objects: c = 8 exp(,0:52 ) (47) where the factor 8 allowed the accumulation into eight discrete bins by taking the integer part of c as the index of the bin (from 0 to 7) (see Section 9.1 and Figure 68). A second criterion for detecting similar problem cases was derived from the photometric reductions and is described in Sections 9.1 and 9.3. In FAST the optical transfer function calibration allowed for the representation of the signal through the three-parameter model, Equation 34. For a true single star transit, the residuals Ik ( ^ ) , Ik (^) should be a zero-mean white noise, with the variance estimated by s2 as in Equation 41. The statistical test used the Fisher ratio F35 : b F35 = p235 =2 s2 =(l , 5) r (48) 288 where: p235 = F. VAN LEEUWEN Xm hIk b , Ik r i (^ ) k=1 (^) 2 Ik (b^ ) (49) was estimated from the residuals of Equation 34. Also, in this case, the Fisher variable 2F35 could be assumed to be asymptotically distributed as a 2 distribution having two degrees of freedom. The null hypothesis (single star) was rejected when F35 > 3:5. The F35 statistic became unreliable for stars of extreme colour, as these could not be covered by the calibration model as described above. 8.4.7. Accuracy assessment At various stages during the software development and testing, first on simulated data and later on actual mission data, comparisons were made between the FAST and NDAC results. These comparisons, led by M.A.C. Perryman, showed initially some problems with the processing of the partially observed stars (see Section 7.2), but after this was solved, any remaining differences were well below the intrinsic noise levels. Both consortia had performed intensive testing before launch using simulated data (see Chapter 2 of Volume III of ESA, 1989). These tests verified the correct implementation of the mathematical model used for the IDT data processing to a high degree of accuracy (0.001 mas for the NDAC tests, using noise free data). Figure 48 shows an example of formal standard errors obtained for the phase and mean intensity estimates for frame transits over part of one orbit. Clearly visible is the effect of variable amounts of observing time assigned to different observations. These data and their error estimates formed the input for the great-circle reduction, the main-mission photometric analysis and the double star analysis. In the greatcircle reductions the data for on average 35 frame transits were combined (all frame transits for a star as obtained during one orbit), while in the photometric reductions on average 9 frame transits were combined into a field of view transit. From these graphs it can be seen that for the majority of frame transits the errors on the phase estimates are between 10 and 60 mas. As attitude errors were highly correlated over short time intervals, the more relevant errors were those averaged over field transits. These were about 3 times smaller: 3 to 20 mas, which, compared with the projection error resulting from errors on the attitude reconstruction, was the dominant noise contributor for all but the brightest stars (see also Section 8.2.1). 8.5. GREAT-CIRCLE REDUCTIONS The scanning of the sky by Hipparcos was such that the z axis moved by about 24 arcmin over one revolution of the satellite. There was, as a result, a considerable amount of overlap between successive full circle scans of the sky. This characteristic led to the concept of the reference great circle and the great-circle reductions. The THE HIPPARCOS MISSION 289 Figure 48. Formal errors on the phase estimate ( 3 , expressed in mas) and the mean intensity ( 1 = 1 ) as a function of mean intensity. The errors are clearly dominated by Poisson noise, shown by the linear relations in the log-log diagram. The discrete line-like features are the result of different amounts of observing time assigned to the measurements at different times, depending on competition. 290 F. VAN LEEUWEN Figure 49. Histogram of the total number of single stars per reference great circle. The full line represents the NDAC data, the dotted line the FAST data. data was accumulated in batches, referred to as data sets. In the original mission concept a data set would have spanned 11 to 12 hours, but due to the orbital problems of the satellite this time limit was effectively set at the total observing time accomplished during an orbital period. Great-circle data sets can therefore be recognized in the final mission products through the orbit number. A reference great circle was associated with each data set. This circle was defined by means of the position of its pole, chosen either as the mean pole position for the data set or as the nominal position of the pole for a time halfway the start and end times of the data set. One data set contained up to 70 000 individual phase estimates, obtained over 4 000 to 16 000 observational frames for between 800 and 2100 different single stars (see Figure 49). The task of the great-circle reductions was to relate the phase estimates obtained from the frame transits for star i in frame k to an actual position on the modulating grid, Gik (using the best a priori positional information for the observed stars as available at the time of the reductions), and to relate this position to an abscissae and ordinate (vi , ri ) with respect to a reference great circle, using the results of the attitude reconstruction. The relation between Gik and (vi , ri ) was described by the instrument parameters, improvements to the star positions and improvements to the along-scan satellite attitude. The satellite’s attitude was provided in three coordinates: two describing the instantaneous position of the z axis, and one describing the scan phase. The first two provided the information on the instantaneous orientation of the modulating grid with respect to the reference great circle, which defined the projection of the observed grid position onto the reference great circle. The scan phase provided the 291 THE HIPPARCOS MISSION provisional linking between the observed transit times and angular displacements along the reference great circle. The phase estimates obtained from the reduction of the IDT data defined a position relative to the modulating grid pattern, without determining where on the grid the image was situated at the reference time. This information was obtained from the a priori positions given in the Input Catalogue and its updates (see Section 8.3). By providing positions with accuracies well below 0.1 arcsec there was generally little or no ambiguity in the identification of the reference slit belonging to a star transit phase estimate. For some 20 000 faint stars these positional improvements were only obtained as a result of the first preliminary sphere solutions. At early stages of the reductions it was not uncommon for the first estimates of the grid position of a star to be wrong by one or more grid steps. These grid-step errors (inconsistent choice of the reference grid line over one great circle) and grid-step ambiguities (consistent but wrong choice of the reference grid line over one great circle) were resolved in the great-circle reductions and sphere solutions respectively. A comprehensive description of the great-circle reduction process was presented by van der Marel (1988) (pre-launch status) and for the post-launch status by van der Marel & Petersen (1992). 8.5.1. The reduction mechanism Using a provisional instrument calibration, the star mapper attitude and the best known catalogue position for an observed star (corrected for proper motion, aberration and relativistic light bending by the Sun and the Earth), predicted values were obtained. The differences between the predictfor the grid positions, Gcalc ik ed and observed grid positions were modelled as a function of corrections to the along-scan attitude (one parameter for every observational frame), the instrument calibration parameters (34 parameters, not all of which were always estimated), and the corrections to the star abscissae (about 1500 parameters): @G @G G = @G v + + d + @v @ @d ik ik ik i i ik k k ik ; (50) where ik is a general noise term including: the photon noise error 0ik ; the projection error on the reference great circle: 00 ik @G @G + ; r + = @G @r @ @ ik i i ik k k ik k k (51) modelling errors in the instrument parameters, the attitude modelling and the star position; the linearization errors The values of Gik were occasionally affected by grid-step errors due to poor quality a priori star positions or due to problems with the attitude reconstruction. These problems largely disappeared with the introduction of better star catalogues. Thus, on average some 50 000 observations equations with 12 000 unknowns were to be solved. This is the kind of problem that is relatively rare in astronomy but 292 F. VAN LEEUWEN fairly common in geodetics. In both consortia, the great-circle reduction software was prepared at or in collaboration with geodetic institutes: for FAST by D. van Daalen and H. van der Marel of the Geodetic Institute of the Technical University in Delft, and for NDAC by C.S. Petersen Copenhagen University Observatory in collaboration with K. Poder of the Geodetic Institute in Copenhagen. The observation equations were written such that the parameters in each of the three groups (stars, attitude and instrument) appeared together: first the very sparse attitude parameters, then the sparse stellar parameters, and last the well covered instrument parameters. Giving the observed grid-coordinate residuals Gik and the unknown corrections to the parameters by the m-vector and the n-vector , the observation equations were written as: y y = Ax + e = Aa xa + Asxs + Aixi + e; x (52) where the subscripts a, s and i refer to attitude, stellar and instrument parameters respectively. The underscores are used to emphasize the stochastic nature of the observations and the residual noise . Given the (diagonal) covariance matrix y of the observations , the least squares solution was computed by solving the normal equations: y e Q x^ (AT Qy,1A)^x = AT Q,y 1y; (53) using Cholesky factorization. After solving and backsubstituting the instrument parameters (see Section 8.5.3), the remaining Cholesky factors were rewritten as two triangular systems, which were solved by simple forward and backward substitution. Even after backsubstitution of the attitude parameters, the stellar part of the normal equations was still very sparse. By storing only non-zero elements and reorganizing the normal equations considerable savings in computing time were obtained. A subset of the non-zero elements of the covariance matrix for the starsolutions was obtained as part of these calculations. This type of solution, where one attitude parameter was solved for every observational frame, was referred to as the geometric solution. The smoothness of the attitude of the satellite allowed, however, the attitude corrections in consecutive observational frames to be linked. 8.5.2. Attitude smoothing and accuracy assessment In the smoothed attitude solution the individual frame parameters a were replaced by a functional representation, using cubic splines, thus reducing considerably the number of unknowns. At times of thruster firings, a knot was introduced with a discontinuity in the first derivative. Other knots were distributed over a thruster firing interval according to the requirements of the data. The smoothed solution was computed as an update to the geometric solution, using the already well determined instrument parameters. The smoothed solution produced much reduced noise levels on the stellar abscissae (see Figure 50). x 293 st.dev. g/c abscissa (smoothing) [mas] THE HIPPARCOS MISSION st.dev. g/c abscissa (geometric) [mas] Figure 50. Star-abscissa improvements by attitude smoothing (data from 21 May 1990, day 506). Figure 51. The abscissa residuals correlation coefficient as a function of the separation on the reference great circle for all data sets. The NDAC curve can be distinguished from the FAST curve by its lower minima and higher maxima, differences that become more pronounced as data sets become shorter. The attitude updates, linked to the positional updates, introduced correlations between the abscissae for neighbouring stars and for stars separated by integer multiples of 58 , the basic angle between the two fields of view. These correlations, 294 F. VAN LEEUWEN Figure 52. Standard deviations ( in mas) for the abscissa residuals as a function of the Hp magnitude. Open symbols refer to FAST data, filled symbols to NDAC data. The diagonal line represents the expected relation for Poisson noise on the photon counts only, given one and the same integration time for all intensities. The flattening of the relation towards brighter magnitudes shows the 2 mas noise contribution from the instrument parameters and attitude modelling contributions. which were estimated before the start of the mission by Lindegren (1988), are shown averaged over all mission results in Figure 51. They tended to be stronger for data sets containing smaller amounts of data (see van Leeuwen & Evans, 1997 and Section 11, where it is also described how to account for such correlations when combining proper motion or parallax data for stars in a small field on the sky). Various comparisons were made between NDAC and FAST results for the greatcircle reductions. One of the prominent features was some instability with respect to a 6th harmonic, causing a modulation of the abscissa results. Such modulations, which could not be suppressed in the great-circle reduction process, were removed in the sphere solution (see Section 8.6). The sensitivity to these harmonics increased with decreasing length of the data sets and with the presence of gaps in the data (resulting from repeated occultations), leading to not fully closed great circles. Occultations were most frequent for great circles having a small inclination with respect to the ecliptic plane. The position of the pole of every great circle is stored in hip rgc.dat on disk 5 of the ASCII CD-ROM set. The final standard errors of the abscissa residuals were the result of two main error contributors: the Poisson noise of the original photon counts and errors in the instrument and attitude modelling. Figure 52 shows the observed standard errors for FAST and NDAC. The better performance of the FAST results for the intermediate brightness range is due to the inclusion of the second harmonic phase information. This gain is becomes less for the fainter stars and causes additional modelling problems for the brighter stars. Towards the brighter stars the errors are dominated by contributions from the instrument and attitude modelling, giving a standard error of close to 2 mas. 295 THE HIPPARCOS MISSION The diagonal line in Figure 52 indicates the expected relation for Poisson noise only. The actual relation deviates from this due to the increasing amounts of observing time assigned to fainter stars. As a result, the signal intensity is not directly proportional to the total number of photons received for an observation, and the relation between standard errors and magnitudes for fainter stars reflects both observing time and Poisson noise. 8.5.3. Instrument parameters The instrument parameters described the relations between an ideal grid and the actual grid for both fields of view, as well as the basic angle between the two fields of view. The instrument parameters evolved over the mission as a result of the satellite and payload settling down to the conditions in space. This was most clearly seen in the settings for the optimal focus. The instrument model consisted of a 2-dimensional third-order polynomial in positional coordinates, the basic angle and chromaticity components for the basic angle, scale and grid orientation. Fourth order terms were in some cases also included. Thus, there was a basic set of 25 instrument parameters, extended to 34 when fourth order terms were included. In the NDAC representation the parameters for the two fields of view were represented by the mean values (g -parameters) and half the differences (h-parameters) between the fields of view (thus, gij hij gives the relevant parameter ij for the preceding field of view, and gij , hij similarly for the following field of view). The parameters were given a two-digit index, indicating the polynomial degree for the x and y coordinate on the grid. Thus, h00 represents half the basic angle (shown in Figure 8), g10 the mean scale along the scan-direction and h10 is half the difference in scale for the two fields of view. g01 represents the mean orientation of the grid and h01 half the difference in orientation for the two fields of view. In the FAST representation the instrument parameters were given per field of view, with the same two-digit index as for the NDAC parameters. The FAST parameters for the preceding field of view are given as apij and for the following + field of view as afij , supplemented by 0 for the basic angle correction at the centre of the main grid. Figures 53 and 54 show the evolution of the first order parameters over the Hipparcos mission. The scale parameter was clearly affected by refocussing (discontinuities), while the orientation was affected by problems with the thermal control electronics. Both parameters show the results of the thermal changes of Sun-pointing mode observations. Also visible, and indicated in Figure 53, are the results of a number of anomalies, such as failures of the thermal control electronics (see also Section 6 and Figure 18). These show as a sudden drift in parameters, followed by a restoration of the normal proceedings at the time the problem was detected and cured. Differences between the FAST and NDAC instrument parameters occurred because of the differences in the definition of the reference phase, with the NDAC 296 F. VAN LEEUWEN Figure 53. Variations in the first order parameters for the FAST instrument model. Discontinuities are in general the result of refocusing, but could also be caused by thermal control problems in the payload. Known problems are indicated in the second graph. Table V provides further information on these anomalies. THE HIPPARCOS MISSION Figure 54. As the previous Figure, but for the NDAC data. 297 298 F. VAN LEEUWEN Figure 55. Definition of the angles and directions used to calculate the relative abscissae on a reference great circle. s is the direction to a star, p the position of the pole of the reference great circle, r the direction of the crossing of the reference great circle with the equator, and v the abscissa value gives the ordinate for this star. for the star at s. The angle 0:5 , reference phase depending only on the first harmonic, while FAST used both the first and the second harmonics (see Section 8.4.1). The instrument parameters were also examined in terms of physical parameters of the optical system by Lindegren et al. (1992b) for the first 12 months of the mission. An examination of all mission results is presented in Section 10.5 in Volume 2 of ESA (1997). This model showed the first order positional changes for the optical elements of the Hipparcos telescope that could be derived from the instrument parameters. 8.6. SPHERE SOLUTION The product obtained from the great-circle reduction was a set of 2341 reference great circles with very precise relative abscissae for on average 1500 stars each. These abscissae had been corrected for aberration and relativistic light bending effects, so the only remaining information consisted of the astrometric parameters: the position as a function of time and the parallax. The sphere solution process was designed to determine consistent zero-points for all reference great circles, so that the very precise relative abscissae could be compared and used to calculate the astrometric parameters. The zero-points were defined by the crossing of the reference great circle and either the ecliptic plane (FAST) or the equatorial plane (NDAC) (see Figure 55). The sphere reconstruction process also took care of the removal of instrument parameters that were out of reach of the great-circle reduction process (chromaticity) and parameters related to instabilities in the great- THE HIPPARCOS MISSION 299 circle reductions (the 6th harmonic). Reports on the pre-launch status of the sphere solution were presented by Lindegren & Söderhjelm (1985), Tommasini Montanari et al. (1985), and in Chapters 7, 8 and 9 in Volume III of ESA (1989). A full report on the implementation of the sphere solution in the data reductions is provided in Chapter 11 of Volume 3 of ESA (1997). Intermediate results on the sphere solutions were presented by Kovalevsky et al. (1995) and in Chapter 16 of Volume 3 of ESA (1997). In the sphere solution three sets of parameters were to be estimated: Corrections to the astrometric parameters of the stars used as reference points; Corrections to the assumed zero-points of the reference great circles; Global parameters. Using a catalogue of a priori astrometric parameters, predicted abscissae were calculated for a selection of primary reference stars, all of which had positions reliable enough to avoid any slit errors and ambiguities to disturb the processing. The differences between the predicted and measured abscissae were collected in a vector . The sphere solution modelled those differences as functions of corrections to the assumed astrometric parameters, and the global parameters , giving the corrections to the zero-points, the 6th harmonic for each reference great circle and the chromaticity in each field of view: v a b Aa + Bb = v + ; (54) A where the vector represents the normalised measurement errors. The matrix contains the derivatives of the abscissae with respect to the astrometric parameters and are described in more detail in Section 8.7. The full set of normal equations is given by: 0 AA 0 B a A0 v A B0A B0 B b = B0 v : (55) a Through elimination of the stellar unknowns , two sets of equations were obtained: i h 0 B B , B0 A(A0 A),1A0 B b = B0v , B0 A(A0A),1 A0v (56) and A0Aa = A0 v , A0 Bb: AA (57) Given Np primary reference stars, the matrix 0 becomes a 5Np 5Np blockdiagonal matrix, i.e. zero everywhere except for Np blocks of size 5 5 along the diagonal. Thus, it becomes a simple process to compute the following two vectors: a A0 A,1 A0 v ; (58) a provided the first estimate of the astrometric parameter corrections, and: where f f v = v , A0 fa: (59) f = 300 F. VAN LEEUWEN Substituting Equations 58 and 59 into Equation 56, a set of equations was obtained for the first approximation of the parameters : h i B0 B , B0 A(A0 A),1A0 B b = B0fv; b (60) where the left-hand side contains a practically filled symmetric matrix of size Nb Nb . After sorting the observation equations on star number, Equation 58 could be solved on a star-by-star basis, giving the provisionally corrected abscissa results of Equation 59. In the process of calculating these provisional results, the information for Equation 60 was accumulated, providing as soon as all stars had been processed. These equations were solved by means of Cholesky factorization. Using the first estimates of the abscissae zero-points and global parameters , the abscissa residuals were recalculated and the process was repeated. The iteration continued until corrections to the parameters became negligible, typically contributing less than 0.001 mas. With this convergence, also the astrometric parameters had reached their final values. This was the NDAC approach. The FAST approach differed in a number of details, in particular with respect to the selection of reference stars, and can be found in Chapter 11 of Volume 3 of ESA (1997). In the NDAC approach, the ‘pre-adjustment’ of the astrometric parameters, by means of the provisional updates , had some advantages. It allowed any well determined astrometric solution to contribute to the sphere solution and it allowed the early detection of outliers and grid-step errors. One problem in the determination of the great-circle zero-points was that the solution had a small rank-deficiency, which did not allow for independent determination of all zero-points as well as the astrometric parameters of the stars. This problem was partly solved by adopting a minimum-norm solution, which effectively defined the average correction for all the zero-points to be zero. The errors on the zero-points were of the order of 0.1 mas were defined as the standard errors on the ‘centred’ variables: identical to the minimum norm values (see Chapter 11 in Volume 3 of ESA, 1997). The rigid but free-floating reference frame in positions and proper motions, as based on the Hipparcos measurements only, was later fixed to a ground-based reference frame (see Section 8.7.4). b b b a 8.7. MERGING AND ASTROMETRIC PARAMETER DETERMINATION The comparisons between sphere solutions and astrometric parameter determinations as obtained by the two consortia showed that there would be a substantial gain to be made in merging the two solutions rather than choosing a single one. The results obtained in the two data reduction chains showed an average correlation coefficient of 0.7 between the remaining abscissa residuals, and merging the two solutions would reduce the uncorrelated errors that resulted from differences in instrument and attitude modelling. In addition, the combined consortia results provided more independent abscissa data than each consortium alone. 301 THE HIPPARCOS MISSION 1.1 Unit-weight error of abscissae FAST NDAC 1.0 0.9 200 400 600 800 1000 1200 1400 1600 Days from 1989 January 0 Figure 56. Evolution of the unit weight errors of the abscissa residuals, based on the formal abscissa errors supplied by the reduction consortia. Each curve is a running median over 50 orbits ( 22 days). 8.7.1. Determination of the variances and correlations In order to preserve the proper statistical information on the final astrometric solutions (the covariance matrices) it was necessary to merge the data at the level of abscissa results, i.e. the great-circle reduction results corrected for zero-points, chromaticity and other global parameters. The basic ideas for the merging of the abscissa results were proposed by C.A. Murray and implemented by F. Arenou at Meudon. During the mission a number of sphere solutions were carried out, including or followed by the determination of the astrometric parameters. The final sphere solutions and astrometric parameter determinations from both consortia were used in the merging process. It was essential to first reduce the two sets to one common reference frame through purely orthogonal rotations. These rotations were at a level of 30 to 50 mas, and do not have any physical meaning. They only reflect the convergence of the working catalogue used in each consortium. The rotations that were applied to the final NDAC and FAST catalogues (referred to as N37.5 and F37.3 respectively) aligned these catalogues with an earlier merged catalogue, referred to as H30. The abscissa records for the N37.5 catalogue did not include data for many of the double and multiple stars, but this type of data (for which the abscissa data are disturbed) was included in the abscissa records for the F37.3 catalogue. 302 F. VAN LEEUWEN 1.2 Unit-weight error of abscissae FAST NDAC 1.1 1.0 0.9 0.8 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Standard errors of abscissae (mas) Figure 57. Unit weight errors of abscissa residuals as functions of the formal standard errors. After applying the rotations, averaged (preliminary) astrometric parameters were determined for stars in common to the two solutions. Abscissa residuals for both catalogues were determined relative to these averaged solutions. New standard error estimates for the astrometric solutions were obtained for both sets. These errors were compared with the standard errors of the abscissae and examined as a function of time over the mission (see Figure 56) and as a function of the abscissa error estimates provided by the consortia (Figure 57). A scaling of the error estimates was introduced that reproduced the proper unit weight errors. The total number of abscissae thus obtained for single stars is shown in Figure 58. The two peaks and the wing towards large numbers of observations are the same features as observed for the photometric data in Figure 15. A spatially resolved graph of these data can be found on page 329 of Volume 1 of ESA (1997), Figure 3.2.3. The correlations between abscissa residuals for the two solutions were modelled as a function of the stellar magnitude, the quality of the great-circle reduction results and the time in the mission. They were obtained by accumulating abscissa residuals for single stars, in common to both solutions. The correlations between the FAST and NDAC residuals were derived from the observed variance of the differences THE HIPPARCOS MISSION 303 Figure 58. Histogram of the number of abscissa data per star for all single stars. As both FAST and NDAC abscissae were counted, even numbers of abscissae are more common than odd numbers, which explains the saw-tooth shaped distribution. between the FAST and NDAC abscissa residuals, deviations F and N : = F + N , 2Var(vF , vN ) : 2 vF , vN , and the standard 2 F N (61) 8.7.2. Fitting models to the abscissa residuals The determination of the astrometric parameters was split into two different problems: the stars with undisturbed abscissa measurements (single stars and double stars with separations less than 0.1 arcsec or a magnitude difference of more than 4 mag), and stars with disturbed abscissa measurements (double and multiple stars and stars in planetary nebulae). The abscissa results could only be used for the first group. Stars in the second group were treated in the double stars analysis (see Section 9.1). The abscissae for the first group, which will be referred to as single stars even though in many cases the stars are not really single, can be fitted with a range of models. Each model is described by a set of parameters a and the changes to the abscissa residuals v resulting from changes in these parameters: v + = X i @v @ai ai; (62) where the vector represents the standard errors on the abscissae. The vector v contains observations from both reduction consortia. Equation 62 is solved by 304 F. VAN LEEUWEN minimizing the 2 value, taking into account the correlations that exist between the consortia data: 2 v , @@va a 0V,1 v , @@va a ; (63) where V is a block-diagonal matrix, i.e. zero everywhere except for blocks of size 2 2 along the diagonal wherever there are FAST and NDAC measurements for the same abscissa value. These data, which are referred to as the Intermediate Astrometric Data, can be accessed on the ASCII CD-ROM, disk 5, and are described as the hip i.dat file in Section 2.8 of Volume 1 of ESA (1997). Their use is also described by van Leeuwen & Evans (1997), from which the following simplification for solving these equations is obtained. The matrix V,1 can be written as the product of a lower triangular matrix and its transverse: ) = ( ( V,1 = T0T: ) (64) The lower triangular matrix T is referred to as the Cholesky square root of the symmetric matrix V,1 . After substituting Equation 64 into Equation 63 and some reorganization, the following equation is obtained: 0 @v @v 2 T v , @a a T v , @a a : (65) = ( ) ( ) The matrix T removes the correlation between NDAC and FAST determinations of the same abscissa residual. Measurements can be recognized as such by means of the orbit number (field IA1 in file hip i.dat). This involves never more than two observations at one time: ! 1 0 F ; T (66) p,NF p1 = F , 1 2NF N , 1 2NF where NF is the correlation coefficient between the FAST and NDAC determinations for the abscissa residual. This correlation coefficient, as well as the standard errors F and N are given in the Intermediate Astrometric Data file in fields IA9 and IA10, with the abscissa residual given in field IA8. Thus, a pair of uncorrelated and properly weighted observation equations is obtained by first writing out the full observation equations for the FAST and the NDAC determination, and then multiplying these two equations by T as defined in Equation 66. When only a FAST or a NDAC observation is available, matrix T reduces to a single element, giving the inverse of the standard error on that observation. This also applies to the remaining observation if either the FAST or NDAC observation was rejected in the solution. 8.7.3. Astrometric parameter modelling The basic astrometric parameter model for fitting the abscissa residuals provided by the Intermediate Astrometric Data file assumes a linear proper motion in right 305 THE HIPPARCOS MISSION Positional Offset J1992.55 J1992.15 30 J1991.25 J1989.95 40 J1990.35 50 cubic quadratic 20 standard 10 0 -10 G -20 200 400 600 800 1000 1200 1400 1600 Days from 1989 January 0 Figure 59. Schematic representation of the baricentric position of the photocentre, as a function of time, in a -type (acceleration) solution. The three curves show an example of cubic solution (full line) and the related quadratic and standard solutions. At the indicated epochs the quadratic or the cubic term vanishes from the solution. The heavy dot corresponds to the position at the reference epoch J1991.25. G ascension and declination and a parallax, giving a total of 5 parameters: ( cos , , , cos , ). In the Hipparcos and Tycho Catalogues cos is given as : (t) = 0 + t p (t) 0 + t + p (t); + (67) (t) = where t is measured in years from the reference epoch J1991.25, and p (t) and p (t) are the parallax factors in right ascension and declination. For the majority of stars the 5 parameter model was all that was required to fit the abscissa residuals. They are the 100 038 stars for which field H59 in the Hipparcos Catalogue was left blank (not counting the 263 stars for which no solution was obtained). For these stars all astrometric data (the five parameters given above, their standard errors and their covariances) are given in the Hipparcos Catalogue, fields H8 till H28. The coefficients @ v=@ a for the five-parameter solution are given for every observation in the Intermediate Astrometric Data file, as fields IA3 to IA7. The five-parameter solution relative to which the abscissa residuals were determined are given for each star in the file header records, fields IH3 to IH7, which also identifies the type of solution accepted for the final catalogue (field IH8). In the process of fitting the astrometric data, measurements could be rejected. The percentage of rejected data is given in field H29. If too many measurements were rejected or the goodness of fit of the solution was unsatisfactory, a solution within which the proper motion was a linear or second order function of time was 306 F. VAN LEEUWEN attempted. These solutions are referred to as ‘G’-type solutions. They added two or four coefficients to Equation 67: (t) = : : : + 12 (t2 , 0:81)g (t) = ::: + + 1 2 (t , 1:69)tg_ 6 1 2 1 2 (t , 0:81)g + (t , 1:69)tg_ ; 2 6 (68) where g d =dt, g d =dt, g_ d2 =dt2 and g_ d2 =dt2 . The time t is again measured in years from the reference epoch J1991.25. The reference times for the first and second order time dependence coefficients were chosen such that their influence on the values derived for the linear proper motion was minimal (see Figure 59). The significance of the g and g_ terms was determined by means of the statistics Fg and Fg_ , derived from covariance matrix of the solution: 2 g ; Fg = g g g g2 g (69) g g g g and similarly for Fg_ . By requiring Fg > 3:44, the same level of significance (0.27 per cent) was obtained as for two-sided 3 criterion for a normal distribution. All the information for the G-type solutions is contained in the DMSA/G file: 2 , hip dm g.dat on disk 1 of the ASCII CD-ROM set. There were in total 2163 7parameter solutions accepted, and 459 9-parameter solutions. Most of these cases are likely to be associated with orbital motions in astrometric binaries with periods of 10 years and longer. Some cases will turn out to be spurious, which is most probably the case for a 7-parameter solution for one of the Large Magellanic Cloud stars. With the provision of the Intermediate Astrometric Data it always remains possible to redo a 7- or 9-parameter solution with different constraints, or as part of an orbital motion solution. When applying 7- or 9-parameter solutions to the Intermediate Astrometric Data, one or two pairs of additional derivatives have to added: @v @ai+5 = 1 2 @v ; (t , 0:81) 2 @ai = 1 2 @v (t , 1:69) ; 2 @ai+3 i = 1; 2; (70) and @v @ai+7 i = 1; 2: (71) The abscissa residuals for 235 stars were solved in orbital motion solutions, often incorporating ground-based data. For these stars the 5 basic astrometric parameters were supplemented by 7 orbital elements. In most cases only some of these were estimated based on Hipparcos data alone, and often estimates were obtained for orbital parameters from a combination of Hipparcos and ground-based data. Some THE HIPPARCOS MISSION 307 examples of these solutions were presented by Söderhjelm et al. (1997) and Martin & Mignard (1997). The DMSA/O file (hip dm o.dat on disk 1 of the ASCII CDROM set) provides all information on the systems dealt with before the finalization of the data reductions. However, in particular for orbital binaries there is still scope for improvements and some new detections, in particular for short period (less than 3 years) systems, as was shown by Bernstein & Bastian (1995) and Bernstein (1997). The systems that qualify for further investigations are those for which a so-called stochastic solution was obtained for the astrometric parameters. The stochastic solution, indicated by ‘X’ in field H59, was applied to 1561 stars for which there was a clear indication of unmodelled disturbances to the abscissa data. These disturbances, which could be signs of orbital motion, unresolved duplicity or multiplicity, but may also be an indication of disturbing nebulae very close to the star, were ‘modelled’ in the form of an additional ‘cosmic noise’, added in quadrature to the standard errors of the observations. The level of cosmic noise added was such that the unit weight variance of the solution became equal to one. If the amount of ‘cosmic noise’ required was more than 0.1 arcsec, the solution was rejected, and no astrometric parameter solution has been provided in the Hipparcos Catalogue. The original abscissa data for all these stars are available in the Intermediate Astrometric Data file, and can be examined further. The details of the stochastic solutions, i.e. the amount of cosmic noise added, is provided in the hip dm x.dat file on disk 1 of the ASCII CD-ROM set. It should be realized that the goodness-of fit parameter in field H30 has no meaning for such solutions. Finally there was a small group of stars (288) for which the abscissa residuals appeared to be magnitude dependent. This group contains mainly large-amplitude red variable stars. There are two possibilities for stars in this group. A companion may become visible when the main component of the system is at a photometric minimum, or, while the main component is at a minimum, its signal becomes much more vulnerable to incidental superpositions. It was not always possible to distinguish these two types of disturbances. In addition, there were problems with properly recognizing duplicity for very red stars. The variability induced movements of the photocentre were modelled by adding the following parameters to Equation 67: (t) = : : : + (t) = ::: + 100:4(Hptot (t),Hpref ) , 1 100:4(Hptot (t),Hpref ) , 1 D D ; (72) where D = 10,0:4(HpC ,Hpref ) d and D = 10,0:4(HpC ,Hpref ) d . The differences in magnitude and in position could not be separated. An example of a type ‘V’ solution is shown in Figure 60. A statistical test very similar to the one described for the g -parameters in the 7- and 9-parameter solutions was applied to (D , D ). However, the criterion for acceptance was set lower, allowing a larger fraction of spurious systems to enter the catalogue. Thus, any variable star that, on the basis of these measurements, could be suspected to be a double star, has been marked as 308 F. VAN LEEUWEN 50 J1991.25 40 Positional Offset 30 photocentre 20 standard 10 0 -10 V -20 200 400 600 800 1000 1200 1400 1600 Days from 1989 January 0 Figure 60. Schematic representation of the baricentric position of the photocentre, as a function of time, in a ‘V’ type solution (variability induced mover). The photocentre describes an oscillating motion between the binary components, in phase with the variations of total luminosity of the system. The dashed line gives the hypothetical motion of the photocentre had the total magnitude been constant, corresponding to the astrometric parameters given in the main catalogue. such in the final catalogue. All information for the ‘V’ solutions can be found in the hip dm v.dat file on disk 1 of the ASCII CD-ROM set. 8.7.4. Transformation to the ICRS The Hipparcos reference frame as obtained from the sphere solutions, the merging and the determination of the astrometric parameters, was rigid but not attached to inertial reference points. An inertial reference frame at mas level accuracy had been established already in radio astronomy, based on radio positions of extragalactic point sources: the ICRF. The final realisation of the Hipparcos Catalogue within the extragalactic reference system was described by Kovalevsky et al. (1997), and the reader is referred to that paper for details. To summarize, the following individual link solutions were available: VLBI observations, multi-epoch observations of 12 radio-emitting stars obtained between 1984 and 1994 (Lestrade et al., 1995); Observations with Merlin, astrometric observations of 9 radio-emitting stars (Morrison et al., 1997a); Observations with VLA of radio-emitting stars between 1982 and 1995 (Florkowski et al., 1985); Optical positions of compact sources, a Hamburg/USNO program to determine precise optical positions of about 400 to 500 selected compact radio sources THE HIPPARCOS MISSION 309 which display optical counterparts (Ma et al., 1990, Johnston et al., 1995, Zacharias et al., 1995); Observations with the Hubble Space telescope, measurements of angular separations between Hipparcos stars and extragalactic sources known from VLBI observations (Benedict et al., 1992); The Lick proper motion program, proper motions for 149 000 stars from 899 fields of the Northern Proper Motion program, determined relative to galaxies (Klemola et al., 1994); Catalogue of faint stars (KSZ), containing 977 Hipparcos stars (Rybka & Ytsenko, 1997); The Yale/San Juan Southern Proper Motion program, containing 4100 Hipparcos stars in 63 fields (Platais et al., 1995); The Bonn link solution, using large epoch differences to obtain very accurate proper motions (70 to 100 years); Hipparcos stars were measured relative to extragalactic radio sources (Tucholke et al., 1997, Geffert et al., 1997); The Potsdam link solution, based on plates taken with the Tautenburg Schmidt telescope; proper motions for 360 Hipparcos stars in 24 fields (Dick et al., 1987); Comparison of Earth Orientation parameters as determined by optical astrometry and VLBI, providing an indirect link of the Hipparcos coordinate system to the extragalactic coordinate system (Vondrák, 1996, Vondrák et al., 1997). The results of all these methods are summarized in Tables X and XI as residuals relative to the adopted solution for the spin components and the orientation components respectively. In the general synthesis of the accumulated data two independent approaches were used, referred to as Method A (by L. Lindegren) and Method B (by J. Kovalevsky). The main problem was posed by the claimed accuracies of some of the determinations, and the main difference between the two methods used to combine the data was in determining appropriate weights for the different determinations. In Method A a first solution was made using the a priori weights. This showed the presence of some clearly unrealistic accuracy estimates. The weight for the most discrepant solution was halved, and the overall solution was repeated. This process was iterated until the overall goodness-of-fit agreed with the expected value. The most severely down-weighted solutions were the SPM program and the Earth orientation parameters. In Method B the covariance matrix for the solution, using the provided accuracies of the various methods, was obtained. From this covariance matrix were derived (under the assumption that all error distributions were Gaussian) the standard error corrections for each contribution. This was done for the spin and orientation parameters separately, as well as for a combined solution. In the final solution the mean of both methods was retained, with differences between the results for the two methods at a level of 0.15 mas/yr for the spin parameters and at 0.10 mas for the orientation components. The final standard deviations for the spin components 310 F. VAN LEEUWEN Table X Summary of the spin components results as residuals with respect to the adopted solution. Accuracies are given in brackets. Method !x VLBI HST/FGS NPM (Heidelberg) NPM (Yale) KSZ Kiev SPM (blue, ) SPM (blue, ) SPM (visual, ) SPM (visual, ) Bonn plates Potsdam plates EOP Spin components (mas/yr) !y !z ,0.16 (0.30) ,0.17 (0.26) ,0.33 (0.30) ,1.60 (2.87) ,1.92 (1.54) +2.26 (3.42) ,0.77 (0.40) +0.15 (0.40) +0.09 (0.18) ,0.20 (0.18) ,0.27 (0.80) +0.15 (0.60) ,1.07 (0.80) +0.23 (0.13) +0.07 (0.15) +0.44 (0.12) +0.30 (0.12) +0.93 (0.34) +0.22 (0.52) 0.93 (0.28) , +0.50 (0.20) +0.58 (0.08) +0.71 (0.18) +0.76 (0.06) 0.32 (0.25) +0.43 (0.50) 0.32 (0.28) , , 0.00 (0.08) ,0.30 (0.07) +0.17 (0.33) +0.13 (0.48) Table XI Summary of the orientation components results as residuals with respect to the adopted solution. Accuracies are given in brackets. Method VLBI MERLIN VLA Hamburg/USNO HST/FGS EOP x Orientation components (mas) ,0.10 (0.47) +1.41 (2.60) +4.27 (4.70) +3.38 (5.00) 6.10 (2.16) +2.33 (0.88) , y z +0.08 (0.49) +0.16 (0.50) ,0.64 (2.20) +0.51 (2.40) ,3.75 (5.30) ,5.76 (5.20) ,0.06 (4.90) ,9.20 (4.70) ,3.25 (1.49) +5.42 (2.14) +7.80 (0.90) epoch 1900+ 91.3 94.0 86.3 88.5 94.3 85.0 was estimated at 0.25 mas/yr and for the orientation components at 0.6 mas. Further improvements of the extragalactic link are still in progress, using new observations from which more accurate radio proper motions and reference positions are derived (Morrison et al., 1997a). 8.7.5. Accuracy assessment of the astrometric parameters As was explained in Section 5, the amount and distribution of the observations obtained with the Hipparcos satellite was very much determined by the characteristics of the scanning law, and as a result the accuracies and covariance matrices as obtained for the astrometric parameters show a clear dependence on ecliptic 311 THE HIPPARCOS MISSION Table XII Median standard errors in parallax, , for the entries in the Hipparcos Catalogue with associated astrometry, divided according to ecliptic latitude and median magnitude (field H44), expressed in milliarcsec (mas). jj Ecliptic latitude ( , deg) 20–30 30–40 40–50 50–60 Hp 0–10 10–20 <2 2–3 3–4 4–5 5–6 6–7 7–8 8–9 9–10 10–11 11–12 12 0.88 0.88 0.91 0.88 0.88 0.92 1.02 1.21 1.49 2.01 2.98 4.35 0.98 0.81 0.82 0.86 0.85 0.90 1.01 1.19 1.47 1.97 2.96 4.19 0.90 0.77 0.79 0.81 0.82 0.86 0.97 1.16 1.43 1.93 2.81 4.11 0.78 0.76 0.74 0.76 0.75 0.81 0.92 1.12 1.37 1.82 2.74 4.10 0.64 0.68 0.65 0.66 0.67 0.73 0.86 1.07 1.35 1.84 2.73 4.00 all 1.23 1.24 1.20 1.15 1.11 60–70 70–90 0.60 0.56 0.55 0.55 0.56 0.61 0.72 0.89 1.11 1.54 2.26 3.67 0.53 0.49 0.52 0.51 0.52 0.55 0.63 0.77 0.94 1.24 1.88 3.49 0.49 0.48 0.49 0.48 0.49 0.52 0.59 0.71 0.86 1.13 1.73 2.51 0.89 0.77 0.71 latitude. This is shown extensively in Section 3 of Volume 1 of ESA (1997). The accuracies are, naturally, also a function of stellar magnitude. Table XII summarizes the accuracies for the parallax determinations. Similar tables for the positional accuracies at the catalogue reference epoch J1991.25 and for the proper motions are presented in Section 3.2 of Volume 1 of ESA (1997). In general, the accuracies surpassed the aims set for the mission by almost a factor of two. In the case of the parallax measurements this meant that the relative parallax accuracies aimed for before launch were achieved for a 5 to 10 times larger volume of space, thus significantly enhancing the scientific impact of the mission. The fact that this was achieved despite the orbital problems has to be attributed to all parties involved in the manufacturing of the satellite, the operations and the data reductions. The parallax accuracies were verified, using the methods described by Lindegren (1995), from the distribution of small and negative parallaxes, through a Bayesian deconvolution in which the distribution functions for the parallaxes and the errors on the parallaxes are estimated simultaneously under the assumption that no negative parallaxes can exist. The method dates back to Hertzsprung (1952), and has been refined since then by several authors. An application of this method, using the preliminary 30-months sphere solution results, showed the predicted external errors to be within 1 per cent of the observed internal errors, well within the accuracy of the determination, and confirming the very high accuracy (Figure 61) as well as the reliability of the formal errors on the Hipparcos parallax determinations. The distribution of parallaxes for 100 019 solutions of the simple 5-parameter type 312 F. VAN LEEUWEN Figure 61. Histogram of the parallax accuracies for 100 019 stars with 5-parameter solutions. Figure 62. Histogram of the distribution of parallaxes for 100 019 stars with 5-parameter solutions. is shown in Figure 62. The small tail of negative parallaxes seen in this figure confirmed the claimed accuracies (see also Section 10.2). The positional accuracies are very much a function of time, and will deteriorate quite rapidly away from the reference epoch J1991.25. Approximate ‘deterioration factors’ can be calculated from the provided accuracies and covariances in the main catalogue. This factor represents the average increase per year within a magnitude interval of the formal standard errors on the positional coordinates and is closely linked to the average proper motion accuracies. Table XIII and Figure 63 show these 313 THE HIPPARCOS MISSION Figure 63. Average positional deterioration factors for (filled circles) and function of Hp magnitude. (open circles) as a Table XIII Average positional accuracies at reference epoch J1991.25 (in mas) and their deterioration factors (in mas/yr) as function of Hp magnitude. Hp mag <2 2–3 3–4 4–5 5–6 6–7 7–8 8–9 9–10 10–11 11–12 > 12 < > < > mas 0.60 0.58 0.59 0.59 0.58 0.62 0.70 0.84 1.04 1.44 2.12 3.46 D D mas/yr 0.48 0.48 0.48 0.48 0.48 0.51 0.58 0.70 0.87 1.17 1.74 2.83 0.67 0.69 0.68 0.69 0.68 0.73 0.83 1.00 1.26 1.76 2.61 4.34 0.53 0.56 0.57 0.56 0.56 0.60 0.69 0.83 1.04 1.41 2.10 3.45 values as a function of magnitude. For an individual star the error propagation is fully described in Sections 1.5.4 and 1.5.5 of Volume 1 of (ESA, 1997), and algorithms with the implementation of the error propagation are provided on the ASCII CD-ROM as ‘C’ and as ‘Fortran’ programs (pos prop.c and pos prop.f). 314 F. VAN LEEUWEN 8.8. TYCHO ASTROMETRY AND PHOTOMETRY The reduction of the Tycho data (the continuous stream of star mapper photon counts) was at various stages interlinked with the Hipparcos data reductions. The task of the Tycho data reductions was to reduce the star mapper photon counts to a catalogue of astrometric and photometric parameters for all stars that could be detected in these signals, approximately one million. This involved as basic steps the prediction of transits (based on preliminary catalogues), the detection of transits (described in the context of the main-mission reductions in Section 8.1), the recognition of transits and the reduction of the intensity information to magnitudes. The relation between transit times and positions on the sky was finally described by the main-mission attitude reconstruction results, including the improvements obtained with the great-circle reductions (Section 8.5), the calibration of the grid geometry as a function of time, which was obtained from the astrometric parameter determinations, and the great-circle zero-points as provided by the sphere solution. 8.8.1. The prediction process The prediction process aimed at deriving from the available information on: the satellite attitude description, the satellite’s and Earth’s orbit parameters, a catalogue of astrometric parameters for a selection of stars, and instrument parameters describing the star mapper grid geometry, transit times for all objects that could possibly be visible in the star mapper data stream. Initially, the attitude used was the on-board real time attitude determination (see Section 7.4), which was not very accurate, in particular early in the mission. The star catalogue used initially was the Tycho Input Catalogue (TIC), based on a magnitude limited selection of stars contained in the Hubble Space Telescope Guide Star Catalogue, upgraded with any information available from the INCA consortium for the preparation of the Hipparcos catalogue and from SIMBAD for general information (photometric and spectroscopic data). The relatively large uncertainties of the star positions in the TIC and the on-board attitude meant that the errors in the predicted transit times were often fairly large, requiring large search windows, and resulting often in misidentifications. Using the NDAC attitude reconstruction for the first few months of the mission, much improved transit time predictions were obtained, leading to more secure identifications. The results of these runs were accumulated in Strasbourg where they were used to create the Tycho Input Catalogue Revision (TICR) based on transits identified during the first part of the mission. In comparison with TIC, TICR contained far fewer stars (1.26 million instead of 3.15 million), but the positional information for these stars was in most cases very much better than in TIC. This catalogue was used in the final processing of the data, together with the on-ground attitude reconstruction by NDAC, improving considerably the prediction accuracy THE HIPPARCOS MISSION 315 and thereby the detection chances, which was very important for accumulating data on the faintest stars visible with the star mapper detectors. 8.8.2. Identification of transits The process of identification of transits had as task to connect predicted transit times with observed transit times, and to predict and identify any possible disturbances through overlapping transits. It was important in this process to clearly distinguish the origins of an observed transit: the field of view, slit group and upper or lower branch for the inclined slits, and to avoid double identifications of observed transits. Any parasitic transits (transits of other catalogued stars that were close enough to the target transit to have caused a significant astrometric and photometric disturbance) were recorded with the transit information, providing the angular separation and the brightness of the disturbing object. After identification of the predicted transits, there remained a number of unidentified observed transits. Information for such transits was only preserved if they were close to identified transits, and could possibly indicate a companion star for the identified object. A total of 160 serendipity stars and 6600 new components were derived from these data. 8.8.3. Tycho photometry In the Tycho photometric reductions the observed intensities were calibrated as functions of slit group, position along the slits, star colour, and field of view for BT and VT . Only 10 000 stars between mag 4.5 and 9 could be used as calibration standards (see Scales et al., 1992, Großmann et al., 1995). Brighter stars could be affected by saturation, and fainter stars were affected by a bias. This bias was due to the detection statistics for the fainter stars. Given a normal distribution of errors around a mean value that is close to or even below the detection limit, the non-detection of all observations below the detection limit will give the impression that the object is much brighter than it really is. A process called de-censoring, described in Chapter 9 of Volume 4 of ESA (1997), was developed to correct photometric data for faint stars for the non-detections. The de-censoring process is illustrated in Figure 64. The comparison between the reference magnitudes and the de-censored magnitudes is shown in Figure 65. The star mapper detector sensitivity was modelled as a function of the ordinate of the star (the position along the slit group) and the colour, both up to second order, and a cross-term of the ordinate and the colour, a total of 6 coefficients. The zero-point defined the response half-way the slits for a colour index B , V = 0:7. Separate calibrations were made for the two slit groups, two fields of view and the two pass-bands (Großmann et al., 1995). The star mapper photometric reductions in NDAC were done with a very similar model and comparisons were carried out between reduced data at various stages of the data reductions by D.W. Evans (RGO). 316 F. VAN LEEUWEN 800.0 Number of measurements 600.0 400.0 200.0 0.0 -1.0 0.0 log (signal) 1.0 Figure 64. The distribution of the observed transit intensities for standard stars within a narrow range of magnitudes close to the detection limit. The thin drawn line is the expected actual distribution, the dotted line the observed distribution, showing the bias due to the detection cut-off. The thick drawn line is the derived corrected distribution. de-censored VT - on-ground VT 0.5 0.0 -0.5 8.0 9.0 10.0 on-ground VT 11.0 12.0 Figure 65. The comparison between the final de-censored VT magnitudes and the true magnitudes for standard stars, showing that de-censoring has largely removed the bias from the final magnitudes. A small remaining bias was removed before the data were entered in the catalogue. The star mapper detector response changed much less over the mission than the main detector response, as the number of optical elements in the detector chain that could get affected by radiation was much smaller. Typical response levels at the start, middle and end of the mission are summarized in Table XIV. An example of the evolution of the star mapper detector response over the mission is shown 317 THE HIPPARCOS MISSION Figure 66. The star mapper response for the over the mission for three colour indices. V channel, vertical slits and preceding field of view Table XIV Star mapper responses in counts/sec for a magnitude 10 star as function of channel, slit group and field of view, over the mission. The reference colour index is B V 0:7. , = Channel Slits Field B Incl Vert Incl Vert Incl Vert Incl Vert P P F F P P F F B B B V V V V 400 900 1500 days since 1989.0 2200 2970 2350 3060 1640 2325 1550 2180 2100 2930 2250 2920 1560 2270 1470 2070 2050 2840 2200 2870 1530 2200 1440 2040 in Figure 66, a similar figure for a different combination of channel, slit group and field of view can be seen in Figure 10.2 in Volume 2 of ESA (1997). These figures were obtained from the NDAC reductions, and were confirmed by the Tycho reductions. The passbands for the Tycho photometry are shown in Figure 25. The relation between the Tycho and Johnson photometry is shown in detail in Section 1.3, Appendix 4 of Volume 1 of ESA (1997). These relations are a function of luminosity 318 F. VAN LEEUWEN , , Figure 67. The relationship between (B V ) (Johnson) and (BT VT ) for luminosity class V stars with accurate photometry. The solid line shows the mean reddening direction for O-B2 stars. class and are affected by interstellar reddening. Figure 67 shows the relation for luminosity class V stars and the displacement direction resulting from reddening. The Tycho photometric data appear at various places in the Hipparcos and Tycho Catalogues. The Tycho Catalogue gives the complete information on the final BT and VT magnitudes and estimates of their standard errors in fields T32 to T36, including a flag indicating how these data were obtained. For Hipparcos stars, any available Tycho photometry can be found in fields H32 to H36. Epoch photometry is in addition available for a selection of 34 446 brighter stars as file tyc ep.dat on disk 4 of the ASCII CD-ROM set, and for a much larger selection (481 553 stars), through the CDS in Strasbourg. General characteristics of the Tycho astrometric and photometric data are summarized in Table XV. 8.8.4. Tycho astrometry The Tycho astrometry process derived the final positions, proper motions and parallaxes for some 1.05 million stars, using the transit time information, the reconstructed satellite attitude and the astrometric parameters for 100 000 single stars derived in the main Hipparcos data reduction. In the process, the large- and medium-scale distortion characteristics of the star mapper detectors were calibrated by means of the 1000́00 Hipparcos stars. This is analogous to the calibration of the instrument parameters for the main detector in the great-circle reductions (see Høg et al., 1992b, Høg et al., 1995, and Chapter 7 in Volume 4 of ESA, 1997). 319 THE HIPPARCOS MISSION Table XV Summary of the data contained in the Tycho Catalogue. Errors on positions (epoch J1991.25), parallaxes and proper motions are given in mas or mas/yr. V Interval of T Median T , mag V N (Tycho) N (not in HIP) <6.0 >11.0 5.38 6–7.0 6.63 7–8.0 7.62 8–9.0 8.62 9–10.0 9.61 10–11.0 10.58 4553 4 9550 55 27750 3485 78029 36511 211107 182773 515029 506720 205934 205275 4.0 5.3 5.0 6.7 8.6 8.3 12.9 16.4 16.0 27.2 34.3 33.5 39.2 49.6 48.6 0.018 0.014 0.024 0.036 0.027 0.049 0.084 0.064 0.117 0.128 0.122 0.200 Median standard errors in astrometry: Position 1.8 2.6 Parallax 2.5 3.6 Proper motion 2.3 3.3 Median standard errors in photometry (mag): 0.003 0.006 0.010 T 0.003 0.005 0.008 T 0.005 0.008 0.014 T T B V B ,V 11.19 All Tycho astrometric data is presented in the Tycho Catalogue, available on disk 1 of the ASCII CD-ROM set as file tyc main.dat. When considering Tycho astrometric data, and in particular parallaxes and proper motions, it should be realized that intrinsic accuracies are low, especially so for fainter stars. In the case of parallaxes this means the occurrence of many (large) negative values. It is not advised to use Tycho parallaxes on an individual basis, and when using Tycho parallaxes on a statistical basis to always include those negative parallax values, as is also explained in Section 11. The proper motions as provided in the Tycho Catalogue are in the process of being upgraded through the use of the positional information collected in the Astrographic Catalogue (Kuzmin et al., 1997). 8.8.5. Quality assessment An astrometric quality flag for the Tycho data was derived from a combination of various criteria: The maximum standard error on any of the 5 astrometric parameters, max ; The signal-to-noise ratio for the detection of the stars, defined on the basis of the distribution of residuals for the astrometric solution. This distribution consisted in general of a sharp peak, flanked by wings. The ratio between the number of points in the peak (n1 ) and the number of points in the wings (n2 ) was given by: p Fs = (n1 , n2 )= n1 + n2: (73) pNast , where In the limit of a sharp image on a low background, Fs Nast n1 is the number of accepted astrometric observations; 320 F. VAN LEEUWEN Table XVI The astrometric quality Q and associated quantities. N is the number of stars in the Tycho Catalogue of each quality class. med gives the median standard error of a position coordinate at the catalogue epoch J1991.25 (the errors at the mean epoch of observation of any given star are typically 5 per cent smaller). Q max Fs (mas) 1 2 3 4 5 6 7 8 9 <5 , 10 , 25 , 50 , 150 5 10 25 50 < 150 < 150 < 150 200 – obs N (mas) >5 >5 >5 >5 >5 >5 3 3 < 300 < 300 < 300 < 300 < 300 300 , 5 < 300 , 5 300 – – – – 23147 70945 259695 430182 146520 41695 37821 28949 13077 6301 med (mas) Astrometric quality 2.6 5.5 13 26 39 44 45 54 – – very high very high high high medium perhaps non-single low perhaps non-stellar low, ‘R’ in Field T42 unassigned, ‘H’ in Field T42 The formal standard error of a single observation, obs , which is a measure of the observed half-width of the image: 2 2 2 obs = 0:25(x + y )(Nast , 5); (74) a half-width of more than 300 mas was generally associated with duplicity; Table XVI shows the relation between the quality parameter Q, given in field T40 of the Tycho Catalogue, and these criteria. 9. Parallel processes in the data reductions The data reduction processes described in this section were handled (semi)independently from the main stream processing described in the preceding section. The results of two of these processes were, however, incorporated directly or indirectly in the final data reductions for the main catalogue: the double star reduction, with its capabilities to recognize likely double stars, also indicated which stars could safely be regarded as single stars. The results of the photometric processing were used in the double-star data reductions, and in case of the star mapper photometry, as colour information in a wide range of instrument calibrations. Other processes described in this section are the treatment of data for solar system objects and the variability analysis of the main-mission photometric data. 9.1. DOUBLE AND MULTIPLE STARS The possibilities to detect and resolve double stars with an astrometric satellite were first put forward by Lindegren (1979a). This was followed by several papers THE HIPPARCOS MISSION 321 Table XVII The double star working group Name Consortium F. Mignard (chair) S. Söderhjelm L. Lindegren M. Badiali H.-H. Bernstein J. Dommanget P. Lampens R. Pannunzio FAST NDAC NDAC FAST FAST INCA INCA FAST on detection possibilities by Vaccari & Delaney (1983), Boriello & Delaney (1985), Söderhjelm (1987), and on techniques for resolving the double star information by Delaney (1983), Delaney (1985), Farilla (1985), Lindegren (1987b), Delaney (1987), Söderhjelm et al. (1992), Mignard et al. (1992a), Pannunzio et al. (1992), Mignard et al. (1995), Gazengel et al. (1995). A detailed account of the final procedures used in the double star analysis was presented in Chapter 13 of Volume 3 of ESA (1997). The double star processing and the preparation of the merged results for the final catalogue were supervised by the Double Star Working Group, of which the members are given in Table XVII. The analysis of the double, and in particular the multiple star information was among the most difficult tasks in the Hipparcos reductions. Although the problem was well defined for double systems with separations above 0.1 arcsec and with magnitude differences below 3 to 4 mag, near to these limits more than one solution for the data was often possible. The information available for the double star processing consisted of three parts: A priori information as collected in the CCDM (Dommanget & Nys, 1994), further enhanced by a data base on speckle interferometric measurements set up in Torino (see Figure 74) and by CCD observations for selected objects; statistical indicators for duplicity, as obtained from the main mission modulated signals; the measured modulation parameters, calibrated for response variations over the grid. In the following sections the detection mechanisms, signal processing, quality assessment and merging are briefly described. For detailed descriptions the reader is referred to Chapter 13 of Volume 3 of ESA (1997). 9.1.1. Detection of double and multiple systems There were three independent indicators for duplicity: two associated with the phase difference and amplitude ratios between the first and second harmonics (in 322 F. VAN LEEUWEN Figure 68. Histograms of the accumulated 2 values for the double star HIP 55 (solid line) and the single star HIP 48 (dotted line), showing the effect of duplicity as an asymmetry in the distribution of 2 values for the three-parameter solutions (see Equations 46 and 47). NDAC combined to a single 2 estimate and accumulated in a histogram, see Section 8.4.6), and one relating the amplitude of the first harmonic to the mean intensity of the signal, reflected in the difference between the dc and ac magnitudes (Section 9.3). These data were available for all 118 000 stars that were observed. Figure 68 shows an example of accumulated 2 data for the double star HIP 55 (separation 3.7 arcsec, magnitude difference 1.7 mag), in comparison with a single star of similar magnitude. Figure 69 shows the ac- and dc-photometric data for the same stars. The ac photometry for a double or multiple star shows systematically fainter magnitudes and has a larger to much larger scatter than the ac magnitudes for a constant star of the same brightness. The effect of duplicity on the modulated signal was a function of the orientation angle of the double star system with respect to the grid lines and the difference between the magnitudes of the components. This showed from a comparison between the observed parameters for the second harmonic in the modulated signal, 4 and 5 in Equation 33, and the values expected on the basis of the OTF calibration (see Section 8.4.5). Two functions, Fc (for the relative amplitude of the second harmonic) and Fs (for the relative phase of the second harmonic) were defined as the ratios between the expected values for single stars and the observed values for double stars of various separations and magnitude differences. Assuming the same colour for the two double star components, Fc and Fs become functions of only the orientation of the pair with respect to the modulating grid, and the magnitude difference, as is shown in Figure 70. The effect of the third criterion, the comparison between the dc and ac magnitudes, is shown in Figure 71. In case all stars were single, the histograms would only show a narrow peak around Hpac , Hpdc = 0. Instead, a surplus of positive residuals is clearly visible in the distributions. These positive residuals indicate weaker signals for the ac than for the dc magnitudes, which is as expected for double and multiple stars for which the modulation amplitude is disturbed. The occurrence of many stars with positive magnitude differences showed that at least approximately 323 THE HIPPARCOS MISSION Figure 69. Epoch-photometry data for the same two stars as used in the previous figure. The top diagram shows the single star HIP 48, the bottom diagram the double star HIP 55. The open circles represent the dc magnitudes, the dots the ac magnitudes, which coincide for single stars, but show large discrepancies for double stars. F5c F4s ∆ m = 0.5 4 3 ∆ m = 0.5 2 3 1 1.5 1 1 2 2 0 1.5 2 -1 1 -2 0 -3 -1 -4 0 60 120 180 240 300 360 0 60 120 180 240 300 360 Figure 70. Variation of the functions Fc and Fs with the projected separation and magnitude 1 2 difference of a double star. The separation is given here as the phase difference between the signals of the individual stellar components; runs from 0 to 360 over the grid period of 1.2074 arcsec. The single star limits for the two functions are one and zero respectively. = , 10 per cent of all stars in the Hipparcos Catalogue could be resolved as double or multiple stars by the main-mission detector. A summary of the detection results in Table XVIII shows, by means of the differences between the FAST and NDAC detections, the difficulties associated with detection close to the separation and 324 F. VAN LEEUWEN Hp Hp Figure 71. Differences between median values for the ac and dc magnitudes. The occurrence of many positive differences is due to the presence of double stars. Detection possibilities vary with magnitude, due to both the accuracy of the photometry and the sensitivity to fainter companions. magnitude difference limits, with in both solutions the occurrence of solved binary systems that were not even detected by the other consortium. 325 THE HIPPARCOS MISSION Table XVIII Number of entries solved or detected only as double stars by each consortium, divided into solved, detected or undetected entries for the other consortium. NDAC Solved FAST Detected Undetected Total Solved Detected Undetected 12710 500 2310 1030 210 160 2250 2190 - 15990 2900 2470 Total 15520 1400 4440 21360 Q ✰ Secondary Celestial Pole Scanning circle ρ θ γ Primary ✰ Gs Gp ky he s Grid nt ed o ject pro Figure 72. Geometry of the observation of a binary star with separation and position angle . The angular coordinate along the scanning circle is denoted G. The signal observed depends only on the projected phase difference between the phase of the secondary (at GS ) and that of the primary (at GP . The orientation of the scanning circle is indicated by . ) 9.1.2. Resolving double stars The fundamental data from which the double star parameters were derived are the modulation parameters obtained in the IDT data processing as given by Equation 29. For a double or multiple star these parameters describe the combined effect of the instantaneous transit times and magnitudes of the various components. Unlike for a single star, the resolution of the data in transit times and magnitudes for individual stellar components from a single transit is ambiguous. Only the combined information of many transits at different scanning angles can provide this resolution. Two different approaches to extract the component data were used by FAST and NDAC. In FAST, the relative astrometry (separation, orientation and any changes over the mission in these parameters) was solved first, followed at a later stage by the absolute astrometry (the absolute position, system proper motion and parallax). Figure 72 shows the basic double star geometry for the relative astrometry, as used by FAST. In NDAC all phases were first related to absolute measurements on the 326 F. VAN LEEUWEN sky, and then solved for the relative and absolute astrometry together. The data were collected for this purpose in case-history files, which have been made available in a slightly modified form on on disk 6 of the ASCII CD-ROM set for 38 535 Hipparcos Catalogue entries, including all those that are known or suspected to be double or multiple stars as the Transit Data file (hip j.dat). Considering that this file provides the only data available for further analysis, and has been directly derived from the NDAC processing, the main emphasis on the double star analysis here will focus on the NDAC approach. A complete description of the FAST approach can be found in Chapter 13 of Volume 3 of ESA (1997). The first step in the NDAC processing was to relate all modulation information to a calibrated and fixed reference system. This meant application of: the photometric calibrations (to make intensities over the grid and over the mission compatible); the instrument parameters obtained in the great-circle reductions (ensuring that phases could be properly related); the along-scan attitude improvements obtained in the great-circle reductions (to obtain accurate relative phases); the zero-points of the great circles as obtained in the sphere solution (creating absolute phase estimates). The phase difference between the first and second harmonic as determined for single stars in the OTF calibration (Section 8.4.5), were eliminated, and the amplitudes = 0:7100 of the first and second harmonics were transformed to fixed values, M = 0:2485, giving for a unit-intensity point source at the reference position and N the following modulation signal: Ak = 1 + M cos pk + N cos 2pk (75) The reference position was defined in barycentric coordinates in the reference frame that resulted from the final sphere solution. A component of a double star has an offset, as measured in baricentric coordinates, of ( , ) with respect to this reference position. The observed instantaneous offset, which is a function of time, is given by this offset plus the offset caused by the parallax of the star. These offsets cause a phase shift in the modulated signal by , which can be obtained from the observed modulation parameters. Defining three derivatives: fx = d=d , fy = d=d and fp = d=d, all dependencies can be taken into account rigorously: = fx + fy + fp (76) The factors fx , fy and fp are given in the Transit Data file in fields JT3 to JT5. The target position is given in fields JP2 and JP3, and in case of two-or more pointing systems in further fields of the Pointing Record. The observed signal for an object consisting of m components, after applying all calibrations as described above, is given by: Ak = Xm Ii i=1 cos(pk + i ) + N cos 2(pk + i )]; [1 + M (77) THE HIPPARCOS MISSION 327 where Ii = 6200 10,0:4H p are the intensities of the components, expressed in counts per sample. This relates to the (calibrated) observed modulation coefficients through the following relations: b1 = X X X , X X , Ii i b2 = M Ii cos i i b3 = M Ii sin i (78) i b4 = N Ii cos 2i i b5 = N Ii sin 2i i The observed parameters b1 to b5 and their accuracies are given in fields JT6 to JT15 in the Transit Data file. As these parameters were dependent on the assumed colour for the system, and this colour may be wrong, colour correction factors are also provided (see Section 2.9.3 of Volume 1 of ESA, 1997). By writing the phases i as functions of the component positions (i ,i ) and parallaxes (i ), a set of 5 non-linear observations was obtained for each field transit, giving an average of some 500 observations per system. In the analysis of double stars a set of 12 parameters was solved for (see Table XIX). This type of solution was necessary for optical double stars, identified by ‘I’ in field DC3 of the Double and Multiple Systems Annex, file hip dm c.dat on disk 1 of the ASCII CD-ROM set. However, for physical binaries it was possible to reduce these 12 parameters to a smaller subset. In solutions of type ‘L’ the two components were assumed to have the same parallax, and in solutions of type ‘F’ the relative geometry of the system was assumed constant. In order for the least squares solution to converge, a priori coordinates were required with errors much smaller than the grid-period of 1.2074 arcsec. This information was not always available, and in many cases a double iteration was required: one iteration providing a range of a priori positions, and the other iteration solving the non-linear equations. It was not uncommon to find a number of indistinguishable solutions, with possible positions for the secondary component found at intervals equal to the grid period. In many cases, however, the solutions were well defined, in particular when magnitude differences were less than 2.5 mag and separations more than 0.3 arcsec. For less well defined solution, a single star solution was obtained as alternative, and accepted if the 2 for the double star solution was not significantly better. A further complication came from double star systems containing at least one variable star. A number of double star systems are in fact triple star systems, with one of the components being a tight eclipsing binary (van Leeuwen & van Genderen, 1997). The modelling as described above no longer strictly applies 328 F. VAN LEEUWEN Table XIX Astrometric and photometric parameters solved for in the NDAC doublestar analysis. Three different physical models used for double stars resulted from the use of various constraints: without any constraints the solution 0 a solution of type ‘L’ results, and with is of type ‘I’, with u10 u10 u11 u12 0 the solution is of type ‘F’. The parameter u13 applies only to double-pointing systems, independent of solution type. = = = = Parameter Description u1 u2 u3 u4 u5 u6 u7 u8 u9 u10 u11 u12 [u13 magnitude of primary (Hp1 ) offset (1 ) in of primary from reference point offset (1 ) in of primary from reference point offset ( 1 ) in parallax of primary from reference point offset (1 ) in of primary from reference point offset (1 ) in of primary from reference point magnitude difference ( m Hp2 Hp1 ) relative position in of secondary (2 1 X ) relative position in of secondary (2 1 Y ) relative parallax of secondary (2 1 ) relative proper motion in of secondary (2 1 X ) relative proper motion in of secondary (2 1 Y ) correction factor for the attenuation profile ( 1)] _ _ = , , , = , = _,_ = _ _ ,_ = _ ' to these systems, and improvements of the astrometric parameters may still be possible for such systems. Systems containing more than 2 components depended even more on good a priori coordinates. Without, the search mesh became unmanageable. The basic equations remained, however, the same. Another approach for multiple star recognition, through imaging, was presented by Boriello (1987). Similar ideas, applied to double stars, were presented by Quist et al. (1997), from which was obtained Figure 73, showing an image of a binary discovered by Hipparcos (see Section 12.2). 9.1.3. Quality assessment and merging The NDAC processing produced a total of 16 016 resolved double-star solutions (this included some of the systems detected initially only by FAST), as well as solutions for 180 triple or quadruple systems. A quality parameter Q was assigned to each solution, based on the difference in magnitude and the separation of the components, as given by: 8 < m , 20 log10(%=0:138) % < 0:10 arcsec D = : m , 5:5 log10 (%=0:32) 0:10 < % < 0:32 (79) m otherwise The discovery limit for a double star was around D = 3:5 to 4. On the basis of D , the quality parameter Q was set to range from 3 (reliable solution) to 0 (marginal THE HIPPARCOS MISSION 329 Figure 73. An image of a binary discovered by Hipparcos. The two objects (HIP 15719) have a separation on the sky of 0.32 arcsec and a magnitude difference of 2.8. The restoring beam is shown in the lower left-hand corner. (from Quist et al., 1997) solution). This quality flag, together with the quality assessment by FAST, was used to define the overall solution quality as given in field DC5. When a priori information was available, the solutions had little chance to converge wrongly. The speckle interferometry information as collected in Torino helped to resolve and verify the solutions for almost 700 systems (see Figure 74). Comparison exercises took place at various times during the mission, mostly associated with one of the intermediate sphere solutions. These comparisons had shown what problems could be expected in the final merging of the consortia results. In the final merging, some 1800 entries for which the solutions as obtained by FAST and NDAC differed by multiples of the grid period, still had to be resolved. Some 700 of these were in the end reduced to single star solutions, and for the remaining 1100 various criteria were applied to decide which should be regarded as the most reliable. The merging of the remaining solutions, for which general agreement had been reached, proceeded following a simple recipe: a straight mean, giving equal weight to the two solutions, was obtained for both the absolute and the relative astrometry. The different approaches in the double star analysis by FAST and NDAC caused another problem: the covariance matrix for the solutions as given by NDAC included correlations between the relative and absolute astrometry, while in the FAST processing these correlations were by default assumed to be non-existing. In the merging this was accommodated through the introduction of neutral points: posi- 330 F. VAN LEEUWEN δ(ρ cos θ) mas 40 20 0 -20 -40 -40 -20 0 20 40 δ(ρ sin θ) mas Figure 74. Relative positions on the tangent plane on the sky between the position of the secondary with respect to the primary obtained by Hipparcos and speckle interferometry. The difference is defined in the sense Hipparcos minus speckle. tions in the system where there was minimal coupling between the absolute and relative astrometry. These positions could be derived from the covariance matrices as provided by NDAC, and were a function of the separation of the two components and their magnitude difference. Within the double star system, it ranged from the position of the primary for separations above 0.7 arcsec to the position of the photocentre for separations below 0.25 arcsec. The merging was thus done in two steps: the relative astrometric parameters from the two solutions were averaged and the absolute astrometry for the neutral points in the two solutions were averaged. The final solution was the combination of these two averaged solutions. Finally, there were some 1400 cases with discrepant FAST and NDAC solutions to be resolved. 60 of these were resolved by means of dedicated CCD observations by Le Poole and others (see Figure 75). 600 other cases showed much better fits in the FAST solution and were kept as such. Remaining systems were investigated using the NDAC imaging approach, which resolved a further 100 cases. All cases retained as valid double-star solutions, but with an alternative solution from the other consortium, were graded as ‘uncertain’ in fields H61 (Hipparcos Catalogue) and DC5 (Double Star Annex C). Some key parameters from the alternative solution are given in the Notes to the Double and Multiple Systems Annex, file hd notes.dat on disk 1 of the ASCII CD-ROM set. There remains the possibility that for newly discovered systems (field H56=‘H’) with relatively poor solutions (field H61=‘C’) both solutions are wrong by one or more grid steps. In some cases improved solutions were obtained after the official completion of the catalogues. Such solutions are indicated by notes to the main catalogue and double star catalogues. 331 THE HIPPARCOS MISSION 15 12 9 % 6 3 0 -10 -8 -6 -4 -2 0 2 4 6 8 10 Reduced δ(∆m) Figure 75. Histogram of the reduced (normalized) differences of the relative photometry between Hipparcos and the CCD observations obtained with the JKT 1-meter telescope on La Palma. The dotted line is a normal distribution of zero mean and standard deviation 1.4. Table XX Quality assessment of double star systems in the main catalogue. A: good; B: fair; C: poor; D: unreliable. Known New Total A B C D Total 7360 1621 8981 904 590 1494 399 428 827 536 356 892 9199 2995 12194 The impact of the Hipparcos double star discoveries is shown in Figure 76. As could be expected, the largest impact is for the smaller separations, although there are also discoveries among the larger separations, even at not very large magnitude differences. Table XX summarizes the statistics for known and newly discovered systems with respect to the quality parameter as given in field H61 in the Hipparcos Catalogue. The relative number of good quality solutions among known systems is obviously much larger than for newly discovered systems due to the advantages of a priori knowledge and generally more favourable parameters for system resolution. 9.2. SOLAR SYSTEM OBJECTS A selection of solar system objects was measured by the main-mission detector and/or by the Tycho detector. 48 Minor planets and three planetary satellites (J II Europa, S VI Titan and S VIII Iapetus) were observed by the main mission. The selected objects had to be bright enough and small enough (diameter < 1 arcsec) to be observed (Lindegren, 1987a). The data for these objects was prepared for 332 F. VAN LEEUWEN Figure 76. The resolved double star systems in the Hipparcos Catalogue. Top diagram: newly discovered systems, bottom diagram: previously known systems. The dotted line in both diagrams indicates the detection limit D=4 as used by NDAC. publication by D. Hestroffer at ESTEC, in collaboration with F. Mignard (see also Hestroffer et al., 1995). Two problems were known to occur when observing these objects: they can not be regarded as point sources in most cases, which affects the phase estimates and their relation to the stellar positions (see Figure 77); their motion on the sky could be so fast, that individual observational frame estimates had to be preserved 333 THE HIPPARCOS MISSION Figure 77. Theoretical differences between the observed position and the photocentre versus apparent 20 . Lower diameter for a spherical object of uniform brightness viewed with a solar aspect angle curve: positions derived using only the first harmonic (NDAC); upper curve: positions derived using first and second harmonics (FAST). = 500 400 300 200 100 0 10 15 20 25 30 35 Figure 78. Distribution of the solar phase angle of minor planets during the Hipparcos mission. Due to the characteristics of the scanning law, observations only occurred in the vicinity of the quadratures. rather than the frame transits combined into single field transits. The first problem resulted from the effect of the solar aspect angle on the phase estimates. The solar aspect angle varied between 13 and 33 : due to the restrictions of the scanning law, observations only occurred in the vicinity of the quadratures (see Figure 78). Reference phases do not refer to the centre of the object, but the differences are generally too small to be of much concern. 9.2.1. Presentation of the data on solar system objects Each observation presented for a solar system object relates to a transit of that object over the main grid, and is based on suitably averaged data obtained from the frame transits, i.e. taking into account the predicted drift of the object during the 334 F. VAN LEEUWEN Figure 79. Definition of the abscissa correction v for a predicted position ( the catalogue reference position ( 0 , 0 ) and the orientation angle . m, m ), with respect to transit. All Hipparcos observations are one-dimensional transit times. This applies to the solar system objects too. However, the Hipparcos Solar System Objects Astrometric Catalogue, file solar ha.dat, gives for each transit both coordinates, a reference 0 and 0 , fields SHA2 and SHA3. This is no more than a reference position, derived using a recent ephemerides. The actual positional information is in this position, combined with the position angle (field SHA6), which define a great-circle arc corresponding to the actual transit position as observed (see Figure 79). The accuracy of the transit position is given in field SHA7. A similar situation exists for the solar system observations for Tycho (file solar t.dat). All observations are expressed in ICRS coordinates, and to a position as seen from Earth. In addition to the astrometric data, also photometric data are provided, but only for the minor planets. The Jovian and Saturnian moons were too much disturbed by veiling glare from the planets. Photometric values for minor planets were obtained by FAST only, and were an implementation of the photometric reductions for stellar images using a fixed colour index of B , V = 0:5. The ac magnitudes are affected by the non-negligible sizes of these objects. All magnitudes have been presented in the solar hp.dat file, and are complemented by information on aspects that affected the observed magnitudes: the distance Sun-asteroid, the distance Earthasteroid, and the solar phase angle. Figure 80 shows an example of the photometry for the minor planet Papagena (471), corrected for distances and aspect angle, and folded with the rotation period of 7.112 d. 335 THE HIPPARCOS MISSION 7.4 7.2 7 0 2 4 6 Figure 80. Folded light curve for (471) Papagena obtained at epoch correspond to observations made about 3 days later. t 1140. The circled points 9.2.2. Using the solar-system-object data The use of the solar-system-object data is very similar to the methods described for single stars in Section 8.7.3 and was described by van Leeuwen & Evans (1997) and used as such by Morrison et al. (1997b) for the observations of Europa and Titan. It is based on providing estimated positions ( m , m ) based on a reasonably accurate ephemerides. Comparing these positions with the observations (the great circles through the reference positions at angles ) gives the abscissa residuals v as: v = ( m , 0) cos sin + (m , 0 ) cos : (80) These abscissa residuals can be modelled in various ways by expressing corrections to the residuals as corrections in right ascension and declination: (v) = d cos sin + d cos + ; (81) with < 2 > = v . By representing d and d as functions of time or as functions d 1 2 of orbital parameters, systematic corrections to the predicted ephemerides positions 2 P, can be obtained through minimizing i vi,vdvi . Observations from differently i oriented reference great circles (angles ) are needed to solve these equations. 9.3. PHOTOMETRY The photometric reductions aimed at the calibration of the intensity measurements for both main-mission (IDT) and star-mapper data as functions of the field of view, the position in the field of view and the star colour. Like for the main-mission astrometric data, the reduction procedures for the photometric data were developed independently at RGO for NDAC (see Evans et al., 1992) and at CERGA for FAST 336 F. VAN LEEUWEN Table XXI The photometry working group Name Consortium D.W. Evans (chair) F. Mignard V. Großmann M. Grenon F. van Leeuwen M.A.C. Perryman NDAC FAST TDAC INCA NDAC ESTEC by Mignard et al. (1992b). All photometric reductions, comparisons and merging was supervised by the photometry working group, with members from NDAC, FAST, TDAC and INCA (see Table XXI). The photometric reductions were made with respect to a set of some 20 000 standard stars distributed all over the sky. The selection of standard stars was partly based on pre-mission ground-based photometry, and partly on photometry accumulated over the first 18 months of the mission. Examination of this photometry also led to the elimination of a number of variable stars among the initial selection of photometric standards. The calibration values for these stars were representing a photometric reference passband close to the actual passband of the system at the beginning of January 1992, as calibrated by M. Grenon at Geneva Observatory from the preliminary photometric calibration parameters obtained by FAST for the period December 1989 till October 1991. The photometric reductions considered the following aspects: the calibration of the background signal. the calibration of the sensitivity of the detectors over the field of view; the calibration of the colour dependence of the detector chain, representing the differences between the actual passband for the data set and the reference passband; the adjustments of the error estimates; the final corrections for positional and time-dependent systematic discrepancies. Each of these aspects was treated independently by FAST and NDAC. The colour dependence was, however, calibrated once again a posteriori by M. Grenon for both consortia, to take properly into account the effects of ageing of the optics on very red stars. There were two photometric indicators: the mean intensity of the signal and the modulation amplitude (see Section 8.4.1). The first is referred to as the dc photometry, the latter as the ac photometry. The dc photometry was about two times more accurate than the ac photometry, but was affected by background THE HIPPARCOS MISSION 337 contributions for stars fainter than 9th magnitude. The ac photometry was sensitive to duplicity and parasitic transits. Most of the calibration procedures, the exception being the background determination, applied to both the dc and the ac photometry. All published data were calibrated with respect to the final set of standards. There were some essential differences between the FAST and the NDAC calibration mechanisms, described in Chapter 14 of Volume 3 of ESA (1997), but various detailed comparisons during and after the data reductions showed no reasons for concern with respect to the overall results of either reduction scheme (see Section 9.3.4). 9.3.1. Background determinations There were at least four contributors to the background in the IDT signal: Galactic glow (the accumulation of faint stars in the Galactic plane), with peak levels of around 20 counts s,1 ; Zodiacal light (reflection of sunlight by dust in the plane of the ecliptic), with peak levels around 30 counts s,1 ; Various detector related background signals, not much more than a few counts s,1 ; Radiation related background (primarily during Van Allen Belt crossings), very variable, peak levels above 100 counts s,1 . The first two of these could be modelled to an acceptable accuracy, while the third could be solved from the available data. The last contribution could only be incorporated by linking it with the measurements of the background in the star mapper detectors. This required a calibration between background responses in the star mapper and the IDT. A few (very rare) orbits within which the attitude control on-board the satellite did not reach convergence (either for the entire orbit or for a large part of it) were used for this purpose. Attitude convergence problems did not hamper the star mapper background determinations, which were obtained independent of whether a star signal was detected. IDT data was still provided for such orbits, but the measurements contained only occasionally, and by accident, a star signal. In most cases all that was measured was a background signal. Figure 81 shows an example of the measured and modelled background for one such orbit. The estimated background corrections were subtracted in intensity scale from the actual observations, leaving only a constant offset to be estimated by the calibrations. This residual background component was generally very small and had little influence on the final results. More complex models could not be fitted as only the intensities of the faintest stars were affected significantly by background contributions, and very few of these faint stars were measured. A more complex model would have had too many degrees of freedom and could have fitted measurement errors rather than real background contributions. Different models were used for the contributions of the zodiacal and the Galactic light by FAST and NDAC. The contribution from the radiation was only accounted for by NDAC. Because of these differences in approach, and the influence of 338 F. VAN LEEUWEN Figure 81. An example of the IDT measurements for a data set without attitude convergence. As a result, the data points show only the background in the IDT detector. The solid line is the modelling of this background by NDAC. Contributions related to zodiacal light are indicated by below the graph, Galactic plane contributions are indicated by . The reaction on the Van Allen belt radiation is much less for the IDT than for the star mapper, as can be seen from the inset, which shows the star mapper background over the same interval. G Z the background contribution on the magnitudes of the faintest program stars, the individual background determinations from both consortia, given in Hz, were preserved in the Epoch Photometry Annex Extension file (hip ep e.dat) as field HTE3 for FAST and field HTE4 for NDAC. The equivalent stellar intensity in Hz at any time during the mission and for any colour is approximately given by: ,8 10,0 4 0 4658 , (82) where = ,9 143 for data obtained before the interruption around day 1350 (see Table V and Figure 18), and = ,9 20 for data obtained after the interruption. : I a (a+ : (t t0 )+F (t) C +H p ) ; : a : The function F (t), the reference times t and t0 , and the pseudo colour index C are defined in Section 9.3.3 and Tables XXII and XXIII. 9.3.2. Positional dependence of the detector sensitivity Two approaches were used in calibrating the sensitivity of the detector chain as a function of position in the field of view. In both groups residuals obtained from preliminary calibrations were accumulated in grids for the two fields of view. In the FAST approach, these residuals were stored as a maps of corrections to be applied 339 THE HIPPARCOS MISSION Figure 82. The observed distribution of residuals in the preceding field of view for the dc photometry. The contour intervals are at 0.005 mag, with the dotted contour representing the zero-level. The range and represents the entire 0.9 by 0.9 field. The graph is based on data in the coordinates obtained between February 5 and March 14, 1990. G H G H Figure 83. As the previous figure, but after removal of the gradients along the and coordinates, showing clearly the circular symmetry of the detector sensitivity over the field of view. Very similar graphs were obtained for the following field of view and for the ac photometry. to the data before the actual reductions. Maps were created for three colour indices, the two fields of view and the ac and dc components of the photometry. In NDAC the general features of the residual distribution were modelled and entered as parameters in the photometric calibration (see Figures 82 and 83). After removing the gradients in G and H , the remaining general feature was a circular symmetry 340 F. VAN LEEUWEN Figure 84. Time variation of the response in the centre of the field for 8th mag stars of two different colour indexes, showing the effect of ageing of the detector chain through the decrease in response over the mission and the difference in this decrease for stars of different colour. around a fixed position and a variable slope. This characteristic was entered as a set of 6 parameters in the NDAC reductions. After all data had been reduced, final position dependent corrections were determined and applied (see Section 9.3.4). 9.3.3. Colour corrections and ageing The passband for the main mission photometry was changing over the mission due to radiation darkening of the IDT relay optics. The loss of transmission was more severe towards the blue than towards the red (see Figure 84). As a result, the effective passband changed over the mission. This change in passband reflected in the colour coefficients derived in the data reductions. A major problem with these colour coefficients was their behaviour towards the extreme red stars ((V , I )C > 2). Most of these stars are large-amplitude variables and highly unsuitable as calibration standards. A special program was set up in collaboration with the AAVSO and observers at ESO, La Silla, for the observations of a large number of such variables. This allowed the determination of a number of reference points for the detector response at extreme red colours over the mission. It was also found that the (V , I )C index was better suited to represent the passband changes than the (B , V )J index. For further simplification of the the relations, a pseudo colour index C (based on the (V , I )C index) was created, for which the ageing effects were practically linear (see Table XXII and Figure 85). The introduction of the C -index allowed the photometry of stars with V , I > 2:0 to be reliably calibrated within the Hipparcos H p system. The evolution of the colour correction factor in the final reductions 341 THE HIPPARCOS MISSION Figure 85. The pseudo-colour index C as a function of (V Table XXII Definition of the pseudo-colour index C = a + b(V 3 I. d(V I ) for different intervals of V , , Interval V 0:85 , 0 85 ,0 48729 , 2 00 ,1 03936 , 20 0 5152 I < V V a : I I > : : : : : , I) , + c(V C. I) , 2 I) + b c d 0.98554 2.14720 0.592 ,0 31968 ,0 416 ,0 0256 0.592 0.0 0.0 : : : (as determined by M. Grenon) is shown in Figure 86. The differences between the colour coefficients for the two fields of view and for the dc and ac photometry that can be seen in Figure 86 reflect the small differences that existed between the passbands of the two fields of view and the different types of photometry. Any such differences were removed in the data reductions. The functional representation given in Table XXIII allows an accurate reproduction of the colour corrections applied to the data. This is important in two cases: variable stars with significant colour variations and stars reduced with the wrong , colour index. For variable stars an average colour was used (sometimes only based on an estimate), and no account was taken of possible colour variations, as such information was simply not available at the time of the data reductions. The result for a variable star light curve can be an increased noise towards the beginning of the mission. For a star reduced with the wrong colour index, the result is a general drift in the magnitudes over the mission. A number of these cases were detected already before the publication of the data, and improved , indices were entered in the main catalogue in field H42, while the value used in the data reductions was preserved in field H75 (not contained in the printed catalogue, but only on the ASCII CD-ROM V I Hp V I 342 F. VAN LEEUWEN Figure 86. The colour correction factor as derived for the dc (top) and ac (bottom) magnitudes. The preceding field of view is indicated with open squares, the following field of view with crosses. The curves show the polynomial fits given in Table XXIII, columns 3 to 6. file hip main.dat). The correction for an H p magnitude measured at epoch t for the difference between the assumed colour index Cold and the newly determined index Cnew is given by: H pnew = H pold + F (t) ( Cold , Cnew ); where the function F (t) is defined in Table XXIII. (83) 343 THE HIPPARCOS MISSION Table XXIII Definitions of the colour correction factors F (t) expressed as polynomial in t t0 , for dc and ac magnitudes. t and t0 are measured in units of 1000 JD, t0 = 2448:6225 = 1 January 1992. , dc approx Stand.dev. const (t (t (t (t (t (t (t (t , , , , , , , , t0 ) 2 t0 ) 3 t0 ) 4 t0 ) 5 t0 ) 6 t0 ) 7 t0 ) 8 t0 ) dc ac preceding following preceding following 0.0031 0.0012 0.0012 0.0017 0.0017 0.0084 0.00303 0.00873 0.0416 ,0.0389 0.09000 0.07720 0.1317 0.340 1.989 0.553 ,0.0537 ,0.04022 ,0.04098 ,0.0431 ,0.18863 ,0.13772 ,0.2384 ,0.6746 ,0.6735 ,1.9387 0.598 0.1322 ,2.9443 2.018 2.052 ,1.3327 ,0.742 0.0478 0.0338 ,0.8277 ,2.9383 ,3.6662 ,1.5601 9.3.4. Final quality checks and merging All data reductions were done at observational frame level. In the published epoch photometry, however, these were merged into field transit data. A total of about 13 million field transits for 118 000 stars were obtained by RGO and CERGA. A field transit contained up to 10 frame transits, with an average close to 8. Some outlying frame transit data were rejected in this process. Standard deviations were determined on the basis of the spread in the observations, but as the number of observations was generally low, these estimates were affected by a Student’s-t distribution, which in particular created some over estimates of accuracies. The accuracy estimates were corrected through an examination of the unitweight residuals for constant stars as a function of magnitude at a range of estimated accuracies. The observed variances provided the correction factors necessary to adjust the estimated standard errors. These adjustments were essential for the variability investigations, described in Section 9.4. They were determined and applied for the FAST and NDAC data, and later for the merged data. Figure 87 shows the final distribution of the unit weight errors for standard stars in the merged epoch photometry. All NDAC epoch photometry was checked for the possibility of accidental superpositions from the other field of view, using the Tycho TICR Catalogue (Section 8.8). A total of about 100 000 such transits were found with a minimum perturbation of 0.01 mag. All these transits are flagged in the Hipparcos Epoch Photometry Annex Extension file (hip ep e.dat) in fields HTE8 and HHE10, which together provide an entry point to the Coincidence file (hip ep c.dat), in which details of the perturbing object(s) are provided. 344 F. VAN LEEUWEN Figure 87. The observed distribution of unit weight residuals for standard stars, compared with the expected probabilities for a unit variance Gaussian distribution. Figure 88. The correlation coefficients between the FAST and NDAC photometric results as a function of magnitude. In the examination of standard star residuals it was noted that during some data sets discrete changes in the detector sensitivity had taken place. These changes, of the order of a few millimagnitudes, remain unexplained, and have been corrected in the FAST and NDAC data a posteriori. After all corrections had been applied, the data sets were ready for comparisons. Given the relatively straightforward reductions of the photometric data, no large discrepancies were expected. The main emphasis was on the correlations between the residuals for standard stars as obtained by the two groups. These correlations were an indication of the success of the instrument modelling in the calibrations, and of the gain there would be in merging the two data sets. Similar to the astrometric THE HIPPARCOS MISSION 345 data, the errors on the photometric data had two contributors: the photon noise on the original data and an instrument modelling error. The first error was the same for both groups, the second error was largely uncorrelated due to the different reduction methods used. The correlation coefficients were generally high, with the exception of very bright and very faint stars (see Figure 88). The very bright stars are most sensitive to the instrument modelling, and the faintest stars were affected by differences in the background modelling. The high degree of correlation meant that there was little to be gained by complicated weighting systems, and instead the photometric data were merged by taking the mean value of the two determinations available. In case only one determination was available, this determination was entered in the catalogue, suitably flagged. The data were split in two files: the Hipparcos Epoch Photometry Annex (hip ep.dat) and the Hipparcos Epoch Photometry Annex Extension (hip ep e.dat). The first of these contains the more accurate dc photometry plus accuracies, epoch information and a quality flag. The second file adds to this the ac photometry plus accuracy, the FAST and NDAC background determinations, the field of view, the position on the sky of the other field of view and the coincidence index, thus allowing more detailed investigations. 9.3.5. Properties of the photometric data Various aspects of the distribution of observations and the total number of observations were described in connection with the scanning law (Section 5, Figures 15 to 17). The general conclusion about the distribution is, that although the average number of observations is relatively high (110), the distribution of those observations is far from optimal for periodicity research on variable stars. In any investigation involving the Hipparcos epoch photometry it is important to consider the standard errors on the individual data points. Due to characteristics of the observing strategy different amounts of observing time were assigned to observations of the same star at different times of the mission. In addition, different numbers of frame transits were used in creating the field transits. This was due to a variety of reasons, such as interruptions in the data stream, loss of data due to telemetry errors and competition from other stars. There are therefore significant real variations in the accuracies of the final data. The distributions of errors for median magnitudes are shown in Figure 89. The wings towards larger errors on the median magnitudes are the result of variability for the dc components, and variability and duplicity for the ac components. The median magnitudes, the associated standard errors and the standard deviations of the epoch photometry are all available in the Hipparcos Catalogue in fields H44 to H47, also giving the number of photometric measurements available in the Hipparcos Epoch Photometry Annex. 346 F. VAN LEEUWEN Figure 89. The distribution of errors on the medians in three magnitude intervals. The distribution for the dc magnitudes is shown by solid lines, for the ac magnitudes by dotted lines. The accumulation to the right of each histogram represents the variable stars, and for ac magnitudes also the double stars. 9.4. VARIABLE STARS All main-mission photometry was subjected to variability investigations. In the majority of cases a statistical test on the photometric data showed no significant signs of variability, but for those stars that did show such signs various tests for periodicity were made. The variability investigations were done in parallel by THE HIPPARCOS MISSION 347 Figure 90. The observed distribution of 2 values (full line) compared with the distribution expected under the assumption that all stars are constant (dotted line). The lowest probability bin accumulated all observations with probabilities below 10,15 . M. Grenon and L. Eyer at Geneva Observatory, and by M.J. Penston and the author at RGO, with assistance from various other people. Reports on aspects of the treatment of the variable stars were presented by van Leeuwen (1997a), Eyer & Grenon (1997), van Leeuwen et al. (1997a) and van Leeuwen et al. (1997b). 9.4.1. Statistical indicators The Hipparcos photometry was reasonably well suited to derive statistical indicators for variability. Care had to be taken, however, of the varying accuracies of the individual observations of a star. A general variability statistic was entered in the Hipparcos catalogue, field H52, as a one-letter code. This statistic was derived from 2 tests on the histogram of residuals as obtained for every star with respect to the median magnitude. The residuals were binned over a range of 5 in unitweight standard error. The number of bins was set according to the number of observations available, ranging from 9 bins for less than 93 observations to 33 bins for more than 200 observations. For every binned distribution there was an equally binned Gaussian distribution with variance equal to 1. The differences between the observed (ho ) and expected (hc ) distributions, for bins with at least 3 observations, were accumulated in a 2 value (Pearson’s test, Papoulis, 1991): 2 = Xm ho (i) , n hc (i))2 ; n hc (i) i=,m ( (84) where n is the total number of observations for a star. The 2 value was transformed in a probability using the NAG-library function G01ECF (returns the lower or upper 348 F. VAN LEEUWEN Table XXIV The minimum peak-to-peak amplitude for stars to be confirmed as variables. Stars having variability at these levels have an average probability of being constant of less than 0.001. The reference points are given for the average of 110 observations. Hp T Hp T 4.5 5.0 5.5 6.0 6.5 7.0 0.010 0.011 0.012 0.013 0.015 0.018 7.5 8.0 8.5 9.0 9.5 10.0 0.022 0.026 0.031 0.037 0.044 0.052 110 110 Hp T 10.5 11.0 11.5 12.0 12.5 13.0 0.063 0.076 0.098 0.124 0.18 0.36 110 Figure 91. The distribution of amplitudes and estimated errors on amplitudes for the 5542 stars in the unsolved variables file. tail probability for a 2 distribution with real degrees of freedom). Figure 90 shows the observed distribution of probabilities compared with the distribution that would have been obtained if no Hipparcos stars were variable. The difference between these two distributions indicate the presence of some 20 000 variables among the Hipparcos stars, but for many of those the variability was too small to allow further investigations. The probabilities are reflected in the main catalogue through a ‘C’ in field H52 for stars found to have a more than 50 per cent probability of being constant (but note that this probability is very much magnitude dependent due to the increasing noise levels towards fainter stars, see Table XXIV), and by ‘ ’ for stars with a probability of being constant between 0.5 and 10,4 . 349 THE HIPPARCOS MISSION It was also possible to estimate statistically the amplitude of variations for stars that were apparently variable, but for which the amplitude and/or the distribution of the observations did not allow a successful period search. For simplicity, a sinusoidal variation was assumed: mi = m0 + m1 cos i ; (85) where m1 is the amplitude that has to be estimated. The variance of the observations mi is then approximated by: R 2 2 2 2 0 cos d < (mi , m0) >= m1 R 2 (86) = 0:5m21: d 0 The observed variance in the observations also contained the noise from the observational errors. Thus, taking into account the standard errors on the individual measurements i through weights wi i,2 , and defining mi mi , m0 , the amplitudes of variations were estimated as follows: = s P mf = 2 ( i Pmiwwi=)(=n(n,,1)1) , 1 ; i = f 2 1 i (87) where n is the number of observations. The amplitudes of the variations equaled 2m1 and were checked against the amplitudes obtained from the fitting of sinusoidal functions for a large number of small-amplitude periodic stars. No systematic discrepancies were detected. These amplitudes can be found in field U11 of the Variability Annex for Unsolved Variables, file hip va 2.dat on disk 1 of the ASCII CD-ROM set. Figure 91 shows the distribution of amplitudes and their estimated errors as obtained for the 5542 stars in this file. In cases where a significant variability was detected with an amplitude below 0.03 mag, the star was indicated as a possible micro variable, ‘M’, in field H52 of the Hipparcos catalogue. A more empirical approach, ignoring the variation of the individual errors on the measurements and fitting the observed estimated intrinsic dispersion to the expected amplitudes as observed for periodic stars, was used in Geneva (Eyer & Grenon, 1997). As was indicated in Section 9.3.3, errors in the V , I colour index used in the data reductions could cause a drift in the final photometry. All data showing significant drifts were checked by M. Grenon to see whether such an error could have been the cause of it. Stars for which this was likely to be the case are indicated by ‘R’ in field H52. Improved V , I values were introduced for the final catalogue, but the epoch photometry was not changed. One source of apparent variability was in fact caused by the observation technique. For double stars with separations of between 5 and 35 arcsec (approximately) and small magnitude differences, the position of one or both stars with respect to the instantaneous field of view (IFOV) during the observations gave rice to variations in the observed intensities. These variations were caused by the exact positioning e 350 F. VAN LEEUWEN Hp Figure 92. Hipparcos dc photometry for a newly-discovered Algol type eclipsing binary, HIP 270, showing through the unequal spacings of the minima a highly elliptical orbit. This star received the name V397 Cep. of the IFOV with respect to the images. With at least one of the stars situated there where the detector sensitivity changed rapidly, small positional changes led to large intensity variations. The measurements for these stars are completely unsuited for variability investigations. They are indicated by ‘D’ in field H52 in the Hipparcos catalogue. Incidental (bright) outliers could occur as a result of the superposition of a stellar image from the other field of view. These were identified in as far as possible and given in the Hipparcos Epoch Photometry Annex Extension and the corresponding Coincidence file, see also Section 9.3.4. The remaining 10 000 variable stars were further investigated for periodicity in the variations. 9.4.2. Periodicity investigations The Hipparcos photometry was rather poorly suited for periodicity investigations, in particular for periods longer than a few days. The distribution of data gaps and their lengths (see Section 5) caused a duty cycle value as defined by Wilcox & Wilcox (1995) of 0.01 to 0.003, well below what these authors considered acceptable. The difficulties with detecting periodicity in the photometric data for variable stars affected in particular the semi-regular variables, for which all too often (probably) spurious periods were found. A range of different methods was used in the period searches. At the RGO, an extensive literature search led to two methods more suitable to be applied to sparse data: the discrete Fourier analysis as described by Scargle (1982) and Scargle (1989) and the analysis of variances by Schwarzenberg-Czerny (1989). The two methods were used in parallel. The folded light curves resulting from periods detected by both methods were visually inspected before being accepted. In the case of eclipsing binaries, a manual intervention was often required, although measures had been taken to recognize probable eclipsing binaries from the distribution of the 351 THE HIPPARCOS MISSION Table XXV Summary of the variable stars investigations. Number of entries variable or possibly variable Periodic Variables Cepheids Sct & SX Phe Eclipsing binaries Other types Non-periodic and unresolved Not investigated 11597 2712 273 186 917 1238 5542 3343 (8237 new) (970 new) (2 new) (9 new) (343 new) (576 new) (4145 new) (3122 new) magnitude residuals. In such cases, periods as found with the methods mentioned above were doubled and checked for performance too. This was mainly relevant to the analysis of variance method, which did not depend on any assumptions about the shape of the light curve. Periods for some Algol type eclipsing binaries were best detected “manually”. An example is the star HIP 270, which due to its elliptical orbit had rather unevenly spaced minima, as can be seen in Figure 92. Alongside the visual checking of the detected periods, a reference data base for papers on photometric measurements of stars contained in the Hipparcos Catalogue was used. This data base had been developed over a period of three years at the RGO, and was also used to enter notes and remarks from the inspections of the light curves. The literature and the more relevant notes have been made available both in printed form and on the ASCII CD-ROM, disk 1, files hp refs.doc, hp auth.doc and hp notes.doc and in Volume 11 of ESA (1997). The backbone of the data base was formed by the General Catalogue for Variable Stars (Kholopov, 1985) and its updates as provided by the IAU Information Bulletin for Variable Stars. At Geneva Observatory different period finding routines were applied to three groups of variable stars: Fourier analysis for small-amplitude variables (Deeming, 1975), (FerrazMello, 1981); Dispersion minimization for large-amplitude variables (Stellingwerf, 1978); Renson’s method for eclipsing binaries (Lafter & Kinman, 1965), (Renson, 1980). Here too, all results were examined manually before acceptance, but no references to literature were used. The final list of periodic variables, as presented in Volume 11 of ESA (1997) and on ASCII CD-ROM disk 1 as file hip va 1.dat, is based on a wide ranging verification of any discrepant periods obtained with the RGO and the Geneva analysis. In those cases where the final choice was not clear, a possible alternative period is given in the notes. Stars for which the period appeared possible but uncertain have in most cases been entered in the table for unsolved variables, 352 F. VAN LEEUWEN Figure 93. The distribution of periods given in the Hipparcos Catalogue as Hipparcos determinations (solid line) and as available in the literature (dotted line). with the possible period given in the notes. Some of the discovery statistics are summarized in Table XXV. The distribution of periods as obtained from the Hipparcos data is shown in comparison with the periods that were known already in Figure 93. This figure shows the problems for ground-based measurements with detecting periods around 1 day, a problem not existing for Hipparcos and consequently the range of periods where most discoveries were made. It also shows that Hipparcos did not contribute much to periods longer than about 8 days. This was partly due to the distribution of the data, but possibly also to the larger amplitudes of variations that are generally associated with long-period variables. These large-amplitude variables were generally well known already. The concentration of new periods between 0.1 and 0.6 days is due to Scuti stars, small-amplitude red variables and Cep stars. For all periodic variables an optimization of the period was obtained through curve fitting. At the RGO a semi-regular spline fit of every light curve was obtained by repeating the sequences of observations and knots 3 times, and using the central section of the spline fit. This did not always provide a perfectly closed light curve, but provided a curve that was well suited for a period optimization. The optimal period was obtained through minimizing the sum of the squared residuals of the observations Oi with respect to the fitted curve f , as a function of a change to the period and the derivative of the fitted curve. As was shown by (van Leeuwen et al., 1997b), these residuals can be given by: () t , t p @f () ; Oi , f (i ) = , i 0 p~ p~ @ =i (88) 353 THE HIPPARCOS MISSION Figure 94. The accuracies for periods of Cepheid variables as determined from the Hipparcos data alone. The dotted line represents the relation p 10,5 p2 . = ~ where ti is the epoch of observation, t0 an arbitrary reference epoch, p the last estimate of the period and p the correction to this period, such that the new p p. This solution was iterated a few times between the curve period p0 fitting and the period optimization. Along with the determination of the period correction was obtained an estimate of the accuracy of the period. These accuracies are approximately proportional to the squared periods, with the proportionality factor depending on the amplitude and coverage of the light curve (see Figure 94). A different method was used in Geneva, where the curve fitting was done by means of Fourier components. A complete set of fitted light curves for 2712 periodic variables is presented in Volume 12 of ESA (1997). A selection of unsolved light curves is shown in the same volume. All light curves can also be found as encapsulated postscript files on disk 1 of the ASCII CD-ROM set, directory curves/. The assignment of variability types to newly discovered variables was often difficult and dubious due to the lack of complementary information. This led, in particular for various (semi-)regular pulsating stars, to types that may very well turn out to be wrong. A special effort was made by C. Waelkens to classify many of the low amplitude variables. Names were assigned to newly discovered variables whenever the discovery appeared reliable. This was done, as is done for all newly discovered variable stars, at the Sternberg Institute in Moscow, where over 3000 discoveries from Hipparcos were given names. ~ = ~+ 354 F. VAN LEEUWEN 10. Verifications of the Hipparcos and Tycho data The verification of the Hipparcos astrometric and photometric data was in many ways provided through the parallel reductions by the two consortia. This ensured an optimal reduction of the data, but could not eliminate any problems that may be inherent to the data itself. There is currently very little material with which Hipparcos astrometric results can be compared. A number of studies were, nevertheless, carried out. 10.1. POSITIONS AND PROPER MOTIONS The only verification of positional precisions possible was with the milliarcsecaccuracy positions obtained in the radio reference frame, ICRF (see Section 8.7.4). This comparison showed that, after transformations had been applied, the positions in the Hipparcos reference system were fully consistent with those obtained with large interferometric arrays in radio astronomy. Differences between the Hipparcos optical reference frame and its predecessor, the FK5, are, on the other hand, locally as much as 0.1 arcsec. Such differences had been noted before from data obtained with the automated meridian circles (Morrison et al., 1990). The verification of proper motions was more complicated because of the differences in epoch coverage between the Hipparcos observations (3.5 yr) and groundbased observations (> 50 yr). As was pointed out by Wielen (1996) and Wielen et al. (1997), orbital motions on a time-scale of 5 to 25 years can disturb the comparison between the Hipparcos and ground-based determinations. This may be the reason why the differences between the proper motions as given in the FK5 Catalogue and proper motions for the same stars given in the Hipparcos Catalogue show a larger than expected dispersion. Only 1233 FK5 stars could be compared, 302 others showed various signs of duplicity (actual double stars, non linear motions, orbital motions etc). The proper motion comparison clearly shows the effect of the meridian-telescope data on which the FK5 is based: as a function of right ascension there are no systematic differences larger than 1 mas/yr, but as a function of declination (where the input from different instruments becomes relevant) differences reach a level of 3 to 4 mas/yr. Thus, instead of being due to orbital motions, the differences could still be at least partly due to problems in the FK5 proper motions. The use of Hipparcos proper motions in the analysis of the Hyades cluster by Dravins et al. (1997) provided an indirect test of their quality. In the analysis, all cluster star are assumed to share the same space motion. Without making assumptions concerning the radial velocities of the cluster stars, the projection effect of the proper motion and the parallax alone are sufficient to reconstruct the spatial structure of the cluster (see Section 11.3). The constraint of the shared space motion leads to improvements by a factor two of the individual parallax estimates. This shows in a cluster HR-diagram through much reduced noise (see Figure 95). 355 THE HIPPARCOS MISSION (a) 0 MHp [mag] 2 4 6 8 All (195 stars) Distances from parallax 10 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 (b) 0 MHp [mag] 2 4 6 8 All (195 stars) Distances from model fit 10 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 (c) 0 MHp [mag] 2 4 6 8 Sample AB (95 stars) Distances from model fit 10 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 B − V [mag] Figure 95. The HR-diagram for the Hyades, using data for all presumed member stars (a and b), and for a subset of 98 stars with a good fit to a model assuming a constant space velocity for all stars, and excluding spectroscopic binaries (c). In (a) the observed parallaxes were used to calculate the absolute magnitudes, in (b) and (c) the estimated parallaxes from the model fit. (from Dravins et al., 1997) 356 F. VAN LEEUWEN Any inhomogeneity in the proper motions would have shown in a disturbed HRdiagram. Instead, the HR-diagram turns out to be of very small intrinsic width. A comparison between differential high precision proper motions in the Pleiades cluster and the Hipparcos proper motions show a general agreement, but possibly also some problems of the kind described by Wielen et al. (1997). 10.2. PARALLAXES The parallax determinations as provided by the Hipparcos mission are in principle absolute, with no systematic errors expected above 0.1 to 0.2 mas. There are few possibilities to check parallaxes for a reliability at such a high level, most of which were presented by Arenou et al. (1995). One of the few fully reliable methods examines the parallaxes for objects that can safely be assumed to be too far away to give a measurable parallax in the Hipparcos data. Using the methods described by van Leeuwen & Evans (1997), a common parallax was solved for the stars in the Magellanic Clouds, giving the following results: LMC SMC = 0:27 0:30 mas = ,0:44 0:65 mas 31 stars 8 stars The results are consistent with the assumed parallax values for the Magellanic Clouds of around 0.002 mas, but do not constrain systematic errors to a very high level. A comparison with dynamical parallaxes for 369 orbital systems (Dommanget & Nys, 1982) showed a consistency with the Hipparcos determinations of 0.25 mas. Accurate parallaxes for binary stars in the Hyades by Torres et al (1997a), Torres et al (1997b) and Torres et al (1997c), using orbital parameters, show excellent agreement with the Hipparcos results, as was shown by Perryman et al. (1997b). There are two other sources of parallax measurements with similar accuracies as provided by Hipparcos: 88 stars in common with the program of the USNO (Monet et al., 1992), (Harris et al., 1997) and 12 parallaxes obtained with VLBI observations of radio stars, 5 of which concern double stars of various types. The comparisons between these data and the Hipparcos determinations are shown in Figure 96 and 97. In Chapter 20 of Volume 3 of ESA (1997) two more comparisons were made, with cluster star parallaxes and with photometric parallaxes using the Strömgren photometric system. Both indicate a reliability of the Hipparcos parallaxes to better than 0.1 mas, but these methods cannot be considered as entirely independent verifications, and some doubt about the use of cluster stars and their ground-based distance determinations appears to be justified after the discovery of a 0.3 mag discrepancy between the photometric and the Hipparcos parallax for the Pleiades cluster by van Leeuwen & Hansen-Ruiz (1997) and for the Pleiades and other open clusters by Mermilliod et al. (1997). 357 THE HIPPARCOS MISSION 200 180 160 USNO parallaxes (mas) 140 120 100 80 60 40 20 0 0 20 40 60 80 100 120 140 160 180 200 Hipparcos parallaxes (mas) Figure 96. Comparison between Hipparcos and USNO parallaxes. 50 45 40 VLBI parallaxes (mas) 35 30 25 20 15 10 5 0 -5 down-weighted stars in DMSA annex -5 0 5 10 15 20 25 30 35 40 45 50 Hipparcos parallaxes (mas) Figure 97. Comparison between Hipparcos and VLBI parallaxes. Two stars were down-weighted and five stars are members of resolved double or multiple stars. 358 F. VAN LEEUWEN V Hp Figure 98. Comparison of Walraven -band measurements with Hipparcos median magnitudes. Above: the direct comparison of the data (the W band is expressed in log ), and the fitted curve. Below: the residuals left after subtracting the fitted curve. V I 10.3. PHOTOMETRIC DATA Hp Comparisons between the Hipparcos photometry and photometric observations obtained in the Johnson, Geneva and Walraven (Lub & Pel, 1977) systems were made. Unlike the ground-based data, the Hipparcos data is not affected by seasonal effects or limited to one hemisphere. A comparison between ground-based systems was made by Pel (1991) which showed systematic discrepancies for the Geneva photometric system. These discrepancies showed again in the comparison between the Hipparcos and the Geneva photometry. passband of the Hipparcos main mission photometry, Due to the wide relations with the intermediate bands of the Walraven, Strömgren, Geneva and similar systems are complicated when examined in detail, and have to be resolved as functions of luminosity class. Examining only constant stars of luminosity and the Walraven band shows classes III to V, the comparison between systematic differences to be well below a level of 0.001 mag (see Figure 98). This Hp Hp V THE HIPPARCOS MISSION 359 , Figure 99. The residuals VGen VHip for stars with declinations below 10 (bottom) and above 10 (top) as a function of right ascension. is in practice only possible if both the Walraven and the Hipparcos photometric systems are free from systematic errors to at least 0.001 mag. The comparison with the Geneva system showed systematic differences as a function of right ascension, as can be seen in Figure 99. In this comparison the magnitudes were transformed to the Gen passband and are referred Hipparcos to as Hip . The observed differences as a function of right ascension are due to errors at a level of approximately 0.005 mag in the Geneva system, as nothing similar is noted in comparisons of the Hipparcos photometry with the Walraven or with the Johnson UBV system, and similar discrepancies had already been noticed in comparisons between the Geneva and the Strömgren, Walraven and Johnson systems (Pel, 1991). V Hp V 360 F. VAN LEEUWEN , Figure 100. The mean differences Vground VHip as function of the apparent V magnitude from the Geneva photometry (filled symbols) and from the UBV photometry (open symbols). The observed differences for the bright stars were the result of saturation effects. For the handful of brightest stars there is likely to be a problem with the Hipparcos photometry due to the failure of the analogue mode (see Section 7.1). Stars that were supposed to be measured in analogue mode had to be measured in photon counting mode, knowing that the signals would be partly saturated. A comparison between Hipparcos, Johnson and Geneva magnitudes is shown in Figure 100. For the brightest star, Sirius (HIP 32349), the situation is obviously the worst, and the underestimate of the brightness by Hipparcos reaches 44 per cent. 11. On the proper use of the Hipparcos data Although the application of parallax data as obtained by Hipparcos may seem simple and straightforward, this only applies when relative errors are very small (well below 10 per cent, Lutz & Kelker, 1973). The user has to remain constantly aware of the way the parallax data are implemented and how accuracies of astrometric parameters are accounted for. Failing to do so can lead to unreliable results and wrong conclusions. In positional comparisons, such as for the use of plate transformations, the Hipparcos positions and their estimated accuracies have to be properly transferred to the epoch of the plate, considering the full covariance matrix in recalculating the standard errors of the transformed positions. This is in particular the case for nearby stars where also the radial velocity can play a 361 P(πH) THE HIPPARCOS MISSION π 0.0000 0.0025 0.0050 πH 0.0075 0.0100 Figure 101. The distribution of the observational error in the observed parallax for a star at 200 pc. (from Luri, 1997) role in the propagation of the astrometric parameters and their estimated errors. In addition, proper motions for these stars are sometimes non-linear. 11.1. CONSIDERING SINGLE STAR OBSERVATIONS Hipparcos measured parallaxes, from which distances can be derived. To obtain these distances is, however, not as straightforward as it may seem, in particular when the relative error on the parallax is larger than about 5 per cent. The error on the Hipparcos parallax is well determined, but this is not the case for derived quantities (Lutz & Kelker, 1973, Luri & Arenou, 1997, Brown et al., 1997a). The probability density function (pdf) for the parallax measurements can be assumed Gaussian. The pdf for derived quantities is given by the pdf for the parallaxes times the derivative of the parallax with respect to the derived quantity, the Jacobian. In the case of distances, this means multiplication of the pdf for parallaxes by (1=R)2 , where R = 1= . Figures 101 and 102 demonstrate this for the parallax measurement of a star at 200 pc. The first of these figures shows the probability distribution for measuring a parallax H for a star with parallax = 0:005 arcsec. The second plot shows the probability distribution for the distances given a parallax of 0.005 arcsec as measured by Hipparcos. Though the direct conversion would give a distance of 200 pc, the most likely distance associated with this parallax is higher. This effect is also known in the literature as the Lutz-Kelker effect (Lutz & Kelker, 1973) and demonstrates the need to interpret the Hipparcos parallax data in terms of parallaxes and not in terms of distances or distance moduli. It was shown by Brown et al. (1997a) how the estimated value for a distance modulus differs from the calculated value on the basis of the observed parallax (see Figure 103), depending 362 P(R) F. VAN LEEUWEN R0 0 200 400 600 R Figure 102. The probability density function for the distribution of distances given a measured parallax of 0.005 arcsec with a formal standard error of 0.001 arcsec. (from Luri, 1997) 1.5 ∆M 1.0 0.5 0.0 -0.5 π=10 mas π=1 mas π=0.1 mas 0 1 2 3 σπ /π Η Figure 103. The bias in absolute magnitudes as a function of the relative accuracies of the parallaxes. (from Brown et al., 1997a) on accuracy of the measured parallax relative to the actual stellar parallax. It is not possible, however, to correct on an individual basis for a bias, as the bias can only be calculated when the actual parallax of the star is known, in which case the correction is of no interest. Correcting on an individual basis, using the measured parallax, will essentially make results only more unreliable. When relative errors on parallaxes are large, the data should only be used in testing models in the sense that the observed parallax is or is not in agreement with the parallax predicted by the model. Low accuracy individual parallax determinations should not be transformed into measurements of distance or distance modulus. 363 THE HIPPARCOS MISSION 2.5 2.0 σM 1.5 1.0 π=10 mas π=1 mas π=0.1 mas 0.5 0.0 0 1 2 3 σπ /π Η Figure 104. The precision on the computed absolute magnitude as a function of H = . (from Brown et al., 1997a) Error estimates for these derived parameters can become completely meaningless (see Figure 104). An example of the strange results authors can obtain when examining Hipparcos data in terms of distances was encountered in a paper submitted to a refereed journal, in which the Hipparcos parallax measurement of 0:05 0:82 mas for the Cepheid V Cen (HIP 71116, for which the expected distance is approximately 721 pc) was interpreted as a distance of 20 000 pc, which was subsequently used to derive a value for the distance modulus. However, the estimated distance predicts a parallax of 1.4 mas, which is within 2 from the observed parallax value of 0.05 mas. The occurrence of negative parallax values can always be accounted for, as long as expected distances are translated into expected parallax values before comparisons are made. The occurrence of negative parallaxes is a simple statistical effect, and ignoring these or trying to explain them as the result of some other astrophysical phenomenon is in fact a misinterpretation of the data. 11.2. CONSIDERING GROUPS OF STARS There are better possibilities in dealing with lower accuracy parallaxes when considering groups of stars. There are, however, also more pitfalls. The best known of these is the Malmquist bias: the error on the estimation of the mean parallax for a group of stars when their apparent magnitudes are affected by absorption (Malmquist, 1922, Malmquist, 1936). This has its equivalent in an easily made selection bias which will distort a data sample. An obvious example is ‘ignoring negative parallaxes’. By doing so, the distribution of parallax values is truncated and as a result shifted to a higher average value. When examining data purely in terms of parallaxes such a cutoff is unnecessary, unlike when an examination of the data is (wrongly) made in terms of distances or distance moduli, and negative par- 364 F. VAN LEEUWEN Figure 105. The relation between photometric parallaxes, calibrated using Hipparcos parallax data, and the observed parallaxes with their estimated formal errors for RRab and RRc stars. The solid line represents the one-to-one relation. allaxes can no longer be properly accounted for. Only final results, when significant in terms of a parallax measurement, should be translated into related parameters, still noting the often asymmetric error distributions for these derived parameters and the fact that the most probable value for a derived parameter is usually not identical to what is found from a direct mathematical derivation of the derived parameter. As was shown by Lutz & Kelker (1973), no attempts should be made to obtain derived parameters when relative accuracies are worse than 17 per cent. An example of how to express a luminosity calibration in terms of the parallax measurements was presented by van Leeuwen & Evans (1997), using the calibration of the absolute magnitude of RR Lyrae stars as an illustration, and applied by (van Leeuwen et al., 1997c) for the calibration of the Mira PL relation. The basic equation for a luminosity calibration describes the absolute magnitudes for a group of stars as a function of a set of observable parameters: M^ = M (ai ) (89) The relation between the observed magnitude m (corrected for interstellar reddening and measured in the same photometric passband as M ), the calibrated magnitude and the parallax (here expressed in mas) is then given by: ^ = 100: e,0:4605(m,M (a )) : i (90) ^ Thus, a correction to the estimated parallax is in first approximation given by: = ,0:4605^ X @M ai: n i=1 @ai (91) As these relations are non-linear, they have to be iterated through applying and obtaining smaller and smaller corrections for the assumed parameters ai . The THE HIPPARCOS MISSION 365 proper criterion for the solution of the parameters ai is the 2 value of the residuals of the parallax measurements. This 2 value is also an indicator for the quality of the model used for the luminosity calibration. In cases when observable parameters in the model are having relatively large errors, methods beyond simple least squares will have to be used to minimize the variances of not only the parallaxes, but also for these specific parameters. Figure 105 shows graphically the implementation of such a solution for RR Lyrae stars. The bias in estimates as described by Lutz & Kelker (1973) does not enter the definition of the luminosity calibration parameters, as the entire description of the problem is done in parallax space. Only when applying the calibration in order to obtain an absolute magnitude estimate, and comparing this estimate with an individual, measured apparent magnitude, do the bias problems reoccur. The application of the calibration provides an estimate of the parallax with a formal standard error, which translates non-linearly into uncertainties in estimated distance and distance modulus. At this stage, however, the uncertainty will in general be reduced in comparison with the parallax uncertainties for the original single-star data. Only if the relative formal accuracies for the parallaxes obtained from the luminosity calibration are sufficiently small (less than 5 to 10 per cent), can the luminosity calibration be applied directly in magnitude space. A somewhat different method was presented by Luri et al. (1996) and Luri & Arenou (1997). Examples of the implementation of their “LM method” were presented by Arenou & Gómez (1997), Gómez et al. (1997a), Mennessier et al. (1997) and Alvarez et al. (1997). The “LM method” uses all available information for a given star: the parallax, proper motion, radial velocity and apparent magnitude as well as any other available physical parameters. The distributions of these parameters are represented by a probability function, from which the expected value of the distance R for a star is calculated, which is then used to obtain an estimate of the true distance RLM . Monte Carlo simulations on a sample of A0 stars showed a very good performance of this method in comparison with various other methods. It produces for every individual star an unbiased estimate of the distance. The “LM method” segregates stars according to kinematical groups, based on proper motions, radial velocities (when available) and distribution in space. 11.3. POSSIBILITIES WITH THE INTERMEDIATE ASTROMETRIC DATA The Intermediate Astrometric Data file (hip i.dat) contains the data used to derive the astrometric parameters for undisturbed single stellar images. These data allow for a re-examination of the original astrometric measurements in the context of alternative astrometric models, which can involve just one star or groups of stars (van Leeuwen & Evans, 1997). When considering single stars, an alternative solution may include a new orbital solution or other additional parameters in the astrometric solution. 366 F. VAN LEEUWEN The main application for the intermediate astrometric data is for determining common parameters (parallax, space motion) of groups of stars (star clusters, associations, Magellanic Clouds) and in luminosity calibrations. The luminosity calibrations proceed along the same lines as described in the previous section, with the exception that the parallax values for all stars are determined from the original abscissa data rather than taken from the catalogue. The advantage of this is a better recognition of bad data points, and the possibility to eliminate weak and potentially damaging data, like the astrometric data obtained for Miras near minimum light, which can be disturbed much more easily than the other measurements by duplicity. In clusters and associations the problem is to derive the best possible reference parallax and space motion for the system. However, as was shown in Section 8.5, the original abscissa data from which the astrometric parameters for stars in a small area of the sky were formed are correlated. Thus, by going back to the abscissae data, these correlations can be accounted for in the determination of the parameters of the system. The methods and the data needed to remove these correlations were described by van Leeuwen & Evans (1997). The problem in all solutions for groups of stars is to express the abscissa data as functions of the model parameters or corrections thereof. This was already shown for luminosity calibrations before. For a star cluster the relevant parameters are the space velocity and distance for a fiducial reference point in the center of the group. The observed proper motions and parallax (and if relevant and available also the radial velocity), are related to expected proper motions (which will be a function of parallax and position on the sky), and the differences are expressed as functions of corrections to the assumed space velocity and parallax of the reference point. As the centre of an open cluster is in general rather poorly defined, the exact position of the reference point should be of little importance. The position for the cluster centre is given by the vector u0 in pc: 0 cos u0 = 0,1 @ sin 1 A; 0 cos 0 sin 0 0 cos 0 (92) and its derivative by 0 cos u_ 0 = 0,1 @ sin 0 cos 0 sin 0 0 cos 0 , sin cos 0 0 0 , cos , sin 10 1 V 0 0 A @ 0 A ; 0 sin 0 0 cos 0 0 sin 0 (93) where the vector u_ 0 is expressed in arcsec yr,1 , and the radial velocity V0 in parsec yr,1 (obtained from dividing V0 , when expressed in km s,1 , by 4.74074). The vector on the right-hand side will be referred to as the velocity vector v: 0V 1 i i vi = @ A ; i i (94) 367 THE HIPPARCOS MISSION = where the index i refers to a star if i 6 0 and to the centre of the cluster if i The inverse of Equation 93 can easily be derived: = 0. 0 cos cos sin cos sin 1 0 0 0 0 0 cos 0 0 A u_ 0 : v0 = 0 @ , sin 0 (95) , cos 0 sin 0 , sin 0 sin 0 cos 0 The space velocity u_ 0 is assumed constant over the cluster : u_ = u_ 0 for all values i of i. Only in the centre of the cluster is this condition not completely true due to internal motions. These are, however, in general quite small (of the order of 500 to 1000 m s,1 ). The astrometric parameters for a star at position ( i , i ) can be related to the astrometric parameters for the centre of the cluster. The relations for cluster star i are given by Equation 95, through replacing the index 0 by i. But, as i 0 , Equation 93 can be back substituted to obtain relations between the observed velocity vector i and the reference velocity vector 0 : u_ = u_ v = 0 Av0; i v v i where the matrix for columns): (96) A has the following elements (first index for rows, second index A11 = cos cos cos 0 + sin sin 0 1 A21 = sin cos A31 = , cos cos sin 0 + sin cos 0 sin , sin cos 0 A12 = A22 = cos 1 A32 = sin cos 0 A13 = , cos sin cos 0 + cos 0 sin , sin A23 = , sin sin (97) A33 = cos sin sin 0 + cos cos 0 1 where = , 0 and = , 0 . Equations 96 and 97 contain the classic i i i i i i i i i i i i convergent point method for cluster distance determination. The convergent point is found at the ( i , i ) value for which the matrix aligns the velocity vector 0 with the radial velocity direction. As the convergent point has to lie on the great circle through the proper motion of the cluster centre, the variables ( i , i ) are not independent. Thus, a set of two equations with two unknowns can be obtained for the observed proper motions, solving for the angular distance between the direction of the cluster centre and the direction of the convergent point, and V0 0 . Combined with the measured radial velocity V0 , the distance 0 of the cluster centre is obtained. The same equations are at the basis of the study by Dravins et al. (1997), who used them to derive for individual stars in the Hyades cluster the parallaxes and A v 368 F. VAN LEEUWEN radial velocities, based on the assumption that all stars in the cluster share the same space velocity. After having derived the space velocity using the parallax, radial-velocity and proper-motion data for individual members, Equation 95 can be used to derive two equations for the observed proper motion components of a star, with two unknowns, the radial velocity and the parallax of the star (see Figure 95). For a more distant cluster the spread in positional coordinates is small and the approximations made in Equation 97 can be used. In most cases it will also be correct to ignore the projection effects of the radial velocities and the relative distance variations. What is left then is a set of equations that describe the proper motion of star i with respect to the proper motion of the cluster centre: i i + sin sin 0 , sin sin i + 0 0 0 0 (98) Using the Intermediate Astrometric Data it is now possible to solve for the cluster proper motion using all cluster members. The procedure, described in detail by van Leeuwen & Evans (1997), is as follows: define an estimated cluster proper motion as a reference value; correct the abscissa residuals for all member stars for the difference between the measured proper motion and the reference proper motion; describe the corrections to the abscissae residuals for all members as corrections to the proper motion of the cluster centre using Equations 98; correct these observation equations for the correlations between abscissae with small separations contained on the same reference great circle; apply a least squares solution to solve for corrections to the assumed cluster proper motion. In this procedure faulty data points can be eliminated if required. Experience shows that some data points rejected in the single star solutions can be reinstated. This type of solution effectively treats the cluster star observations as if they are observations of a single star, situated in the centre of the cluster. The same procedure can also include an estimate of the parallax of the cluster. This is done in the same way as was done for the common proper motion. The least squares solution provides a 2 value, showing the correctness of the model that was applied (in this case the constant space velocity and common parallax). It is expected and observed that this model is not entirely correct due to the internal motions in the centre of a cluster. It may be possible to combine the Hipparcos solution with the information from the much more precise differential proper motion studies, but this appears to be a far from trivial exercise, with difficulties caused by the much longer time-span used for ground-based proper motions, and distortions in ground-based proper motion reference frames (Tian et al., 1996). To summarize, the Intermediate Astrometric Data offer a range of possibilities, from a re-examination of the data for an individual star, to a proper treatment for THE HIPPARCOS MISSION 369 Figure 106. An example of uv coverage for a Hipparcos object. This object (HIP 15719) was scanned from virtually all angles. The middle point gives the total flux from all scans, while the two circles represent the two spatial frequencies for the Hipparcos sampling. (from Quist et al., 1997) stars in open clusters, associations and the Magellanic Clouds (Kroupa & Bastian, 1997, van Leeuwen & Evans, 1997). 11.4. POSSIBILITIES FOR THE TRANSIT DATA The Transit Data in file hip j.dat make it possible to re-examine signal-modulation data for signs of duplicity. The main areas of application are double stars of relatively small separation or large magnitude difference, double stars containing one variable component, and multiple stars. Most other situations have already satisfactorily been resolved in the standard reductions. This type of investigation requires input parameters: an estimate of the separation(s) and magnitude difference(s) in the system, to be obtained from independent measurements. As was shown by Quist et al. (1997), the Hipparcos main-mission data has features very similar to data obtained with radio interferometers, and similar techniques can be used in examining those data. The two harmonics in the main detector signal produced 3 unique points in the uv plane, equivalent to baselines of 0, 10, and 20 cm, with the centre of the uv plane being equivalent to the zero-baseline, and giving the total flux of the observed object. Figure 106 shows an example of a well scanned star (HIP 15719). 370 F. VAN LEEUWEN Figure 107. The raw image (dirty map) of Hipparcos object HIP 15719. This map is the Fourier transform of the previous figure. (from Quist et al., 1997) An image obtained by the Hipparcos satellite is derived by taking the Fourier transform of the uv plane along with the phase and amplitude information. This produces what is called a dirty map, requiring further image processing. Figure 107 shows this for the same star as used in Figure 106. Using the CLEAN algorithm as contained in the Caltech Difmap package, the original image is restored (see Figure 73). A users package for examination of the Transit Data, describing these data in UV FITS files (Greisen & Harten, 1981) as used in radio-interferometry, is in preparation by C.F. Quist and L. Lindegren (Lund Observatory). 12. First scientific results This section can by no means be a comprehensive review of the scientific results of the Hipparcos mission, the range of results and details is simply too wide and too many studies have only just started. In preparing it, much use was made of the proceedings of the Venice’97 symposium, supplemented by a few papers that have already appeared in refereed journals, or for which preprints were available. 12.1. NEARBY STARS, LOCAL GALACTIC KINEMATICS Using nearby stars the Hipparcos data clarified to some extent the reliability of photometric/spectroscopic distance determinations for stars of different luminosity THE HIPPARCOS MISSION 371 Figure 108. The distribution with respect to distance of 5610 objects selected on spectral type and apparent luminosity to lie within 80 pc. 2384 of these were found to have larger distances. (from Binney et al., 1997) classes, and the kinematics and population of the solar neighbourhood. Studies of this type are providing a better basis for distance calibrations of objects too distant to have their parallax measured, and of the history of the local group in our Galaxy. A first examination of stars contained in the Catalogue of Nearby Stars (Gliese, 1969, Gliese & Jahreiß, 1979, CNS3 as available at the CDS in Strasbourg) confirmed aspects of distance determinations based on photometric and spectroscopic data that had been suspected before by Grenon (1984) on the basis of a Geneva photometric survey of these stars. The Hipparcos parallax data established that some 40 per cent of stars in the CNS3 lie beyond 25 pc, the limit set for this catalogue (Perryman et al., 1995). A significant number was found to have parallaxes of less than 20 mas. This was confirmed as a more general trend by an examination of a selection of stars for which the Michigan Spectral Catalogue (Houk & Cowley, 1975, Houk, 1978, Houk, 1982) indicated a luminosity class V, which, given their apparent brightness, would put these stars within 80 pc (Binney et al., 1997, Murray et al., 1997). A histogram of the Hipparcos distances derived for these stars is shown in Figure 108. A diagram like this should, however, be presented as a histogram giving the numbers of stars as a function of measured parallax rather than derived distance, as was done by Perryman et al. (1995) for the stars within 25 pc. This allows it to be compared with the formal errors on the parallaxes. A careful examination will now be required to find out how the discrepancies between the expected and observed distances for these stars came about, and what relevance they may have with respect to other studies. One field which has depended for a long time on simple associations between spectral features and luminosities is the study of the kinematics and distributions 372 F. VAN LEEUWEN Figure 109. The variation with colour B dispersion S . (from Binney et al., 1997) ,V of the components of solar motion U , V , W and the Figure 110. Velocity dispersion as a function of age. Open symbols: CNS2-values; filled symbols: new results. Circles: main sequence stars; triangles: K and M dwarfs. (from Jahreiß & Wielen, 1997) of stars within 100 to 200 pc from the Sun. New relations between spectral classifications and absolute magnitudes and their dispersions were determined using selections of nearby stars (Houk et al., 1997). This study, however, seems to show signs of using a sample truncated at a fixed parallax value similar to those indicated by (Luri & Arenou, 1997). The overall results confirm to within a few tenths of a magnitude the classical values tabulated by Schmidt-Kaler (1982), but provide in addition estimates of the intrinsic dispersions in absolute magnitudes for different spectral types and luminosity classes. These dispersions were found to be considerable larger than than they were expected to be. Preliminary results of various kinematic studies using Hipparcos data are now available. The mean velocities and velocity dispersions for groups of stars within 80 pc was investigated by Binney et al. (1997) as a function of colour (see THE HIPPARCOS MISSION 373 Figure 111. The dependence of U , V and W on S 2 . The dotted lines correspond to the linear relation fitted (V ) or the mean values (U and W ). (from Binney et al., 1997) Figure 109). The velocity dispersions were represented by a combined value S 2 , which shows a clear relation with stellar colour: from increasing over the blue stars to constant over the red stars. This effect, first noted by Parenago (1950), reflects the varying mean ages of stars of different colours, and the increase in velocity dispersion towards higher ages. Similar effects were derived from the Hipparcos data by Gómez et al. (1997b) and Jahreiß & Wielen (1997), who split up stars according to age-groups (see Figure 110). Gómez et al. (1997b) also claimed that the velocity ellipsoid gets rounder with increasing age. Strömberg’s asymmetric drift equation (Binney & Tremaine, 1987) predicts a linear dependence of the V -component in the velocities (the direction of the rotation in the Galaxy) and S 2 : when the velocity dispersion in a group increases, this group will rotate more slowly around the Galactic centre. This relation is shown clearly by the Hipparcos data (see Figure 111), and can be extrapolated to a zero dispersion, which represents the velocity of the Sun (U0 , V0 , W0 ) with respect to the local standard of rest: the velocity in a closed orbit in the Galactic plane that passes through the location of the Sun. The following values were found for the local standard of rest: U0 V0 W0 = 11 0 0 6km s,1 = 5 3 1 7km s,1 = 7 0 0 6km s,1 : : : : : : (99) : These values are significantly different from the classical values of Delhaye (1965) in U0 and V0 . This may be related to the difficulties in associating the correct luminosity class to some types of stars (see Figure 108). The distribution of nearby stars was used by Crézé et al. (1997) to derive the potential well and local dynamical mass. The availability of distances and proper motions from the Hipparcos Catalogue has now made it possible to trace 374 F. VAN LEEUWEN Figure 112. The stellar luminosity function derived from the stars within 25 pc. The units are the number of stars within 20 pc per one magnitude bin. The dashed histogram shows the LMF from Wielen et al. (1983). (from Jahreiß & Wielen, 1997) the potential well to 125 pc from the Sun, using 2977 A-type stars. The local dynamical density was found to be 0 = 0:076 0:015M pc,3 , close to the figure for known matter of 0:085M pc,3 . Using the observed value, a dynamical model of the Galactic disk was built, showing that there may be just room for a spherical halo of dark matter (0.008 M pc,3 ), but there is no room for a significant amount of dark matter to be concentrated in the disk. The local luminosity function has been investigated by Murray et al. (1997) and Jahreiß & Wielen (1997). The first of these studies looked at stars within 80 pc, the second within 25 pc. Hipparcos discovered 119 new stars within 25 pc and 37 previously unknown companions of nearby stars. The nearest of these is HIP 103039 at 5.5 pc. The completeness of the sample is very different for declinations north of = ,30 than for southern declinations as a result of the Luyten surveys for high proper motion stars, and the fact that high proper motion stars were selected for the Hipparcos Input Catalogue. Figure 112 shows the derived luminosity function. The main differences with earlier determinations are that the density of giants appears to be lower, while the density of the faintest main sequence stars is higher. The estimate of the total density of stars has decreased from 0.045 pc,3 to 0.039 pc,3 . 375 THE HIPPARCOS MISSION 3.2 M/Msun 1.6 0.8 0.4 0.2 0.1 0 2 4 6 8 10 Hp(abs) 12 14 16 Figure 113. The mass-luminosity plot for all masses with an uncertainty below 15 per cent. (from Söderhjelm et al., 1997) 12.2. STELLAR MASSES Further details on the properties of the nearby stars were obtained from mass determinations derived from orbital parameters of binary systems for which the Hipparcos data also provided an accurate parallax measurement (Söderhjelm et al., 1997). A re-examination of the orbital-system data improved in a few cases the parallax determination with respect to the value published in ESA (1997). The sums of the masses were determined for 45 systems, and accurate dynamical component masses for 15 of these. For the remainder the mass ratio was estimated from the magnitude difference. The well determined masses range from 0:36 0:02M to 3:12 0:37M . Only main sequence stars were investigated. Figure 113 shows the relation for these stars between their H p magnitudes and masses. A similar study, but covering astrometric binaries in which the motion of the photocentre for 43 systems were investigated, was presented by Martin & Mignard (1997). Here masses range from 0.3 to 5 M . Many stars contained in Söderhjelm et al. (1997) were also contained in this study, and differences between masses found in the studies are often larger than expected from the formal standard errors. This may at least partly be due to a reexamination of the orbital and parallax data for these stars in both studies, and shows some of the difficulties and uncertainties encountered in interpreting such data. A preliminary study of the mass-luminosity relation by Lampens et al. (1997) indicated a wide range in slopes and an average value somewhat higher (K = 3:9) than it was previously assumed to be. The high level of noise on the observed relations and the differences in results noted above seems to indicate that further improvements to the analysis of these data can still be made. Some extreme cases of astrometric binaries were examined by Bernstein (1997). Examining those stars for which the single star solution showed the presence of 376 F. VAN LEEUWEN He .014 40 Eri B .012 EG 50 .01 Procyon B Fe GD 140 .008 Mg C .006 .004 .3 .4 .5 .6 .7 .8 .9 1 1.1 Figure 114. Observations of white dwarfs compared with various mass-radius relations. (from Provencal et al., 1997) disturbances on the astrometric data, a number of low mass companions were discovered. In one case, for the 0.03 M companion of Gliese 433 (HIP 56528), the companion was “seen” in the results of 2D IR speckle interferometer observations at ESO SHARP on the NTT by Leinert and Woitas (see Bernstein, 1997). The observed separation and position angle were in very good agreement with the values predicted from the Hipparcos data analysis. Using a combination of theoretical relations and Hipparcos distance determinations, mass estimates were obtained for horizontal-branch stars by de Boer et al. (1997a, 1997b). Using spectroscopic determinations of Teff and the surface gravity, log g , combined with the apparent luminosity transformed to an absolute luminosity by means of the Hipparcos parallax determinations, the masses of horizontal-branch stars were found to be MHB = 0:38 0:07M , considerably below what was expected, although a similar result had been obtained for HB-stars in globular clusters (see de Boer et al., 1995 and references therein). The findings of these studies show that the HB-stars in the field and in globular clusters are very similar, but results would become very different (by 20 per cent) if the change in the galactic distance scale as proposed by Feast & Catchpole (1997) is adopted. The difference between the expected masses and the calculated masses is most likely due to the estimates of the surface gravity values, and an increase in these values by 0.2 would bring the calculated masses more in agreement with predicted masses. 377 THE HIPPARCOS MISSION Figure 115. Cumulative distributions for stars in the Hyades cluster versus distance from the centre 1 M (solid line, 68 objects); 1 2 M (dotted line, 97 objects); and 2M for (dashed line, 26 objects). (from Brown et al., 1997b) M <M M> Improvements have also been obtained in the determination of masses and radii for white dwarfs (Vauclair et al., 1997, Provencal et al., 1997). In addition to 3 white dwarfs in binary systems, there are now also data for 7 white dwarfs in common proper motion pairs, for each of which masses and radii were measured. The masses are all in the range 0.45 to 0.75 M , the radii are between 0.011 and 0.015 R . There is some marginal evidence of a confirmation of the mass-radius relation expected for white dwarfs (see Figure 114). Radii were also estimated for 11 field white dwarfs. 12.3. STAR CLUSTERS AND ASSOCIATIONS Members of some 200 open clusters had been selected for the Hipparcos Input Catalogue by J.-C. Mermilliod with the aim of determining cluster distances. The biggest impact of the results of the mission is for the nearest clusters: the Hyades, Pleiades, Praesepe and Per, all of which have now accurately determined distances and well defined main sequence tracks in the HR-diagram. The parallax and proper-motion data for Hyades-cluster stars for the first time allowed a spatial resolution of the cluster (see Perryman et al., 1997b, Brown et al., 1997b, Dravins et al., 1997) and a reconciliation of the various distance determinations made in the past. A total of 218 probable members were selected, of which 39 had not been considered before as members. From this sample, the 378 F. VAN LEEUWEN Figure 116. Distribution of individual parallaxes and their accuracies for the 54 single Hipparcos Pleiades members. (from van Leeuwen & Hansen-Ruiz, 1997) position and space velocity of the cluster centre were calculated. This was done for two selections: considering stars within the tidal radius of 10 pc, and for all stars within 20 pc. The differences between the results are small. They show a distance for the cluster centre of 46:34 0:27 pc, equivalent to a distance modulus of 3:33 0:01 mag, and a velocity of 45:93 0:23 km s,1 . The available proper motion, radial velocity and parallax data show a full consistency for stars within 10 pc from the cluster centre to have a uniform space motion, with an additional velocity dispersion in the cluster centre of 0.3 km s,1 . Beyond the radius of 10 pc there appears to exist a halo of stars no longer bound to the cluster, as was expected for an older open cluster from numerical N-body simulations (Terlevich, 1987). The consistency of the space motion in the cluster was shown by a very narrow main sequence for the Hyades by Dravins et al. (1997) (see Figure 95), who used it to derive more accurate distances for individual cluster members. The spatial resolution of the cluster also made it possible to distinguish very clearly the mass segregation in the cluster (see Figure 115). A subsequent spectral analysis of 20 single stars in the Hyades cluster by Lebreton et al. (1997) gave the [Fe=H] value for the cluster and from this a value of X = 0:024 0:003. ZAMS modelling was used to derive the cluster age (625 50 Myr) and helium abundance of Y = 0:26 0:02 (Cayrel de Strobel et al., 1997). The helium abundance is very close to the solar value, despite the higher than solar metallicity of the cluster. The much larger distance of the Pleiades cluster implied that there was no possibility to do as detailed an investigation as was done for the Hyades. The improvement in the accuracies of the astrometric parameters from the original aims of 2 mas by almost a factor two had a very significant impact on the possibility THE HIPPARCOS MISSION 379 Figure 117. Colour-magnitude diagram for the Hipparcos Pleiades members. Data for 60 stars are shown, 54 as used in the combined astrometric solution and 6 stars with non-single solutions. (from van Leeuwen & Hansen-Ruiz, 1997 to determine a reliable distance for this cluster. Figure 116 shows the distribution of parallaxes and their accuracies for the selected stars, Figure 117 shows the HRdiagram for the selected stars as well as 6 other members that were known to be non-single. In the distance determination the methods as outlined by van Leeuwen & Evans (1997) were used (see also Section 11.3). Combining the intermediate astrometric data for 54 single stars, van Leeuwen & Hansen-Ruiz (1997), and independently Mermilliod et al. (1997), derived a parallax of 8:61 0:23 mas, which, at an accuracy of better than 3 per cent, can be translated into a distance of 116 pc and a distance modulus of 5:32 0:05 mag. In the same process the cluster proper motion was determined to an accuracy of about 0.2 mas/yr, thus enabling the very precise studies of internal motions to be calibrated to absolute motions. The value found for the cluster distance is substantially different from any determination made on the ground using mostly photometric properties of cluster members (130 to 135 pc), and puts the entire main sequence of the Pleiades at the blue envelope of the Hipparcos HR-diagram (based on stars with parallax determinations better than 10 per cent). Although it is in principle possible to fit the 380 F. VAN LEEUWEN Table XXVI Parallax determinations for open clusters by Mermilliod et al, 1997 (left) and improved values by Robichon (Turon, priv.comm.) (right) Cluster N mas mas N mas mas Coma Ber Pleiades IC2602 IC2391 Praesepe Per Blanco 1 26 51 13 7 24 32 11 11.34 8.60 6.81 6.78 5.65 5.43 3.96 0.22 0.24 0.22 0.25 0.31 0.22 0.48 31 61 13 12 24 49 11 11.49 8.54 6.81 6.79 5.57 5.49 3.96 0.21 0.22 0.22 0.22 0.29 0.19 0.48 Pleiades main sequence with a synthetic isochrone, the helium abundance value required seems far too high to be realistic: Y 0:35. No signs of such high helium abundances are observed in other young open clusters (Brown et al., 1986). It appears close to impossible to measure the helium abundance in the Pleiades, as its most massive stars are of type B6III and are known to rotate very rapidly. The metallicity of the cluster stars is supposed to be close to solar, with [Fe=H] = 0:026 (Mermilliod et al., 1997). Without some observational parameter able to explain the observed difference of 0.3 between the expected and the observed distance modulus, further detailed interpretation of the Hipparcos data becomes rather uncertain. It seems very unlikely that the parallax determination is wrong. If this were the case, then all Hipparcos parallaxes could be assumed wrong, including those that confirm earlier estimates. There seems to be no reason why the Pleiades parallax values should be singled out as wrong. Accepting that the parallax is well determined, and that the metallicity of the cluster is indeed close to solar (but new determinations are required, in particular concerning abundances of oxygen and other elements), there is an uncertainty at a level of about 0.3 mag in photometric distance determinations, which may well interfere with photometric determinations of abundances. Until the position of the Pleiades main sequence is properly explained as a function of observational parameters, determinations of the age of the cluster and of the masses of its stars will remain uncertain. Like the Hyades cluster, the HR-diagram for Pleiades cluster is intrinsically of very narrow width. The observed width is, however, affected by differential reddening in the cluster, duplicity and the spread in distance moduli (Hansen-Ruiz & van Leeuwen, 1997). This narrowness of the main sequence must mean that these clusters were formed from very homogeneous, well mixed clouds. A difference exists between the distance determinations for the Praesepe cluster by van Leeuwen & Evans (1997) and Mermilliod et al. (1997), citing values for THE HIPPARCOS MISSION 381 Figure 118. Distances derived from the mean parallaxes of the Hipparcos members of OB associations versus previous best estimates of distances to these associations. (from de Zeeuw et al., 1997) the mean parallax of 5:04 0:36 mas (based on 22 members) and 5:65 0:31 mas (based on 24 members) respectively. This leads to a difference in distance modulus of 0.26 mag, and possibly shows the vulnerability of cluster parallax determinations to the selection of probable members. Further cluster parallax determinations by Mermilliod et al. (1997) are summarized in Table XXVI, together with a more recent update by Robichon (Turon, private comm.). Stars in OB-associations formed a significant part of the Hipparcos Input Catalogue, primarily as a result of the program proposed by the SPECTER group of Leiden Observatory (de Zeeuw et al., 1997). The aims for the Hipparcos data requested by this project were much improved membership determinations, providing in particular information on probable fainter members, and improved information on the kinematics of these groups. The data was supplemented by a wide range of other observations: Walraven photometry, mm and radio observations of the interstellar medium and precise radial and rotational velocities. The method used to select members is described by de Bruijne et al. (1997). It uses the stellar position, proper motion and radial velocity to select a group of stars all moving towards the same converging point. The dispersion of the proper motion components perpendicular to the direction of the convergent point is used to determine a 2 value for the solution, and to detect and reject non-members as outliers. Examples of the application of this method are shown by de Bruijne et al. (1997), Hoogerwerf et al. (1997) and de Zeeuw et al. (1997). Figure 118 shows a comparison between Hipparcos and pre-Hipparcos distance determinations for 11 associations. The general trend is for Hipparcos distances to be smaller than previous ground-based estimates. 382 F. VAN LEEUWEN 12.4. LUMINOSITY CALIBRATIONS As a result of the higher than expected parallax accuracies obtained by Hipparcos, it became possible to obtain the first direct, parallax based calibrations of the period-luminosity (PL) relations of Scuti stars, Cepheids and Miras. The best possibilities for a PL relation calibration were provided by the Sct stars, and in particular for a small selection with high amplitudes such as SX Phe, AI Vel and V474 Mon (Høg & Petersen, 1997). It was shown that these stars are most likely main sequence stars, probably related to a similar type of variable occurring in globular clusters among the so-called blue straglers. For these stars it is possible to distinguish fundamental mode and overtone pulsation periods (and many show both), and to relate these periods to masses and a period-luminosity relation, for which (Høg & Petersen, 1997) give: MV = ,3:74 log P0 , 1:91; (100) where P0 is the period for the fundamental mode. This relation was, however, derived in magnitude space (using 6 stars), where it is susceptible to bias due to nonlinear transformations (see Section 11.2). A reevaluation by means of Equation 91, using the accuracies on the parallax measurements as weights, gave the following result: , MV = ( 3:98 0:52) log(P0 =0:1) + (1:80 0:09); (101) with a unit weight standard error of 1.16, and using 5 stars with fundamental mode periods: V474 Mon (HIP 28321), AI Vel (HIP 40330), V703 Sco (HIP 86650), RS Gru (HIP 107231) and SX Phe (HIP 117254), and VZ Cen (HIP 42594), for which a fundamental mode period of 0.2263 d was used (the observed first overtone period divided by 0.77). At the reference period of 0.1 d the covariance between the slope and zero-point solutions is only 0.14. An apparently similar star, AD CMi (HIP 38473), lies very much outside the PL relation, as was noted by Høg & Petersen (1997). The comparison between the predicted and measured parallaxes using this relation is shown in Figure 119. A similar (pre-Hipparcos) relation for SX Phe stars in fundamental mode pulsation given by Nemec & Lutz (1993) is very similar but has larger uncertainties: ,(4:19 0:70) log P0 , 1:71 + 0:32[Fe=H]: (102) With [Fe=H] = ,1:70 for SX Phe, this gives a zero-point equivalent to 1:94 in MV = the Hipparcos relation (at P0 =0.1 d). The selection of stars used in the calibration appears from the proper motions to be a mixture of Population I and II stars, with SX Phe and possibly also RS Gru being Population II, and the remainder Population I. There were 21 stars classified as SX Phe in the Hipparcos Catalogue, primarily on the basis of the amplitude and period of their variations, offering still further scope for checking and improving their PL relation. THE HIPPARCOS MISSION 383 Figure 119. Comparison between calibrated (C) and Hipparcos (H) parallaxes for SX Phe stars. Figure 120. HR diagram for a sample of 75 Scuti stars. The continuous lines are the ZAMS and the upper limit of the main sequence, while the dotted lines are evolutionary tracks for the indicated masses and metallicity of Z 0:020. (from North et al., 1997) = 384 F. VAN LEEUWEN Figure 121. The positions in the period versus amplitude diagram for three groups of stars following the classification in the Hipparcos Catalogue. Open circles: Sct; filled circles: SX Phe; open stars: RRc; filled stars: RRab and RR. Also shown, with filled triangles, are periodic variables of spectral types A and F in the same range of periods, but without types assigned in the Hipparcos Catalogue. Most of the stars with low amplitude and periods longer than 0.5 d are Ap stars (see Figure 124). The PL relation for low-amplitude Sct stars is much less well defined due to the mixture of pulsation modes for these stars, as was shown by North et al. (1997). The luminosities derived from the Hipparcos parallaxes and Strömgren photometry show a wide spread, with some 15 per cent lying above the theoretical upper limit of the ZAMS (see Figure 120). The available theoretical models and Hipparcos luminosities indicate a mixture of pulsations in fundamental mode, first and third overtone. The luminosities obtained from photometric calibrations for these stars are, according to Liu et al. (1997), now found to be on average too low, although a considerable scatter is shown to exist, with discrepancies of over 0.3 mag. This may well be related to a similar discrepancy observed for the Pleiades cluster by van Leeuwen & Hansen-Ruiz (1997). The RR Lyrae stars did not provide serious possibilities for a parallax based calibration of their absolute magnitude, while a kinematical calibration did not change much with respect to similar ground based calibrations (Fernley et al., 1997). Only for RR Lyrae itself was a marginally useful parallax obtained: 4:38 0:59 mas, which, following Brown et al. (1997a), gives an uncertainty of about 0.26 mag on the absolute magnitude, and a possible bias at a level of about 0.01 mag. The relation given by Fernley et al. (1997) for the RR Lyrae magnitudes is: MV = (0:18 0:03)([Fe=H] + 1:52) + (0:72 0:04): (103) THE HIPPARCOS MISSION 385 It is interesting to notice that the relation for the SX Phe stars, when extrapolated to a period of 0.4 d, gives an absolute magnitude within a few hundredth of the absolute magnitude of the RR Lyrae stars. In the Hipparcos variability annexes, the distinction for A and F type stars with periods in the range 0.05 to 0.8 d, between Sct, SX Phe and RRc stars was not always clear from the Hipparcos photometry only, as is demonstrated by Figure 121. Classifications for all these stars need to be reinvestigated with the aid of additional ground-based information, and supplemented with the kinematical evidence provided by Hipparcos. A group of small-amplitude periodic variables with periods from 0.5 to 25.0 d also seen in diagram primarily consists of peculiar A stars with almost one quarter having Ap Si as spectral classification (see also next section). The Cepheids, being intrinsically much brighter than the RR Lyrae stars, contained a larger number of significant or marginally significant parallax determinations which were used by Feast & Catchpole (1997) for a calibration of the zero point in the PL relation: hMV i = log P + ; (104) where P is the photometric period. The parallaxes for a selection of 23 stars were used in a solution of the type described by Equation 91, in which the zero point was described as a function of the observed parallax, photometric period and reddening corrected V magnitude: 100:2 = 0:01100:2 hV i, log P : ( 0 ) (105) The Hipparcos data does not contain sufficient information to calibrate the slope of the PL relation, which was fixed at = ,2:81, a value derived for 88 Cepheids in the LMC by Caldwell & Laney (1991). Two types of solution were made, one assuming an intrinsic scatter in the Hipparcos solution of 0.1, and one assuming no such intrinsic scatter. The results from the two solutions are very similar, and a final value of = ,1:43 0:10 was adopted. This is equivalent to a value of 100:2 = 0:517 0:025, giving an uncertainty in parallax values derived from this relation of approximately 5 per cent. At this accuracy level the bias from the non-linear transformations to distances and distance moduli is much smaller than the uncertainties in the calibration itself. Thus, the Hipparcos data were interpreted as parallax data, at which stage there was no bias entering the descriptions. Only at the last stage, when a derived parallax value was transformed to a distance modulus value could a bias as described by Lutz & Kelker (1973) become relevant, but the relative accuracy of the derived parallax was such that this bias is insignificant. When applied to the LMC Cepheids, the calibration of gives a distance modulus for the LMC of 18:70 0:10, rather different from what is derived using RR Lyrae stars or the ring around SN1987A. A similar calibration of the PL relation for Miras showed results compatible with the Cepheids (van Leeuwen et al., 1997c, Whitelock et al., 1997). A calibration of oxygen-rich Miras using the “LM method” 386 F. VAN LEEUWEN 6.0 7.5 -5 8.0 Mv 8.5 0 9.0 9.5 5 Supergiants-Bright giants Giants 1 Giants 2 Clump giants Old giants Main sequence Subgiants -1.0 -0.5 0.0 0.5 (B-V)0 1.0 1.5 2.0 Figure 122. Groups resulting from the “LM-Method” calibration for G and K stars. Isochrones with log(age) are also shown. (from Gómez et al., 1997a) by Alvarez et al. (1997) showed the need of considering different populations when examining the Miras. A number of more general luminosity calibrations have already been started, most noticeably by Gómez et al. (1997a). Bright stars with luminosity classifica- 387 THE HIPPARCOS MISSION 8.0 Mv (Phot) 7.0 a) 6.0 5.0 4.0 3.0 2.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 Mv (Hip) Mv (Phot) − Mv(Hip) 1.0 b) 0.5 0.0 −0.5 −1.0 −3.0 −2.0 −1.0 0.0 1.0 [Fe/H] Figure 123. The new photometric calibration of absolute magnitudes by Jordi et al. (1997). Full circles are stars with [Fe=H] < 0:5, and open circles stars with [Fe=H] > 0:5. (from Jordi et al., 1997) , , tions were subjected to the “LM method” described by Luri et al. (1996), Arenou & Gómez (1997) and Luri & Arenou (1997) (see also Section 11.2). The “LM method” split the stars in different groups according to kinematic properties and spatial distribution and showed clearly a much larger range in absolute magnitudes than was assumed before, with a range as large as 2 mag for especially earlier type stars. It also showed the difficulties in obtaining fully reliable luminosity classifications from spectral properties, with some stars classified as luminosity class V found near the horizontal giant branch and vice versa. This is the same problem that was encountered in the selection of some nearby stars on spectral and luminosity characteristics (see Section 12.1). Figure 122 shows how the LM calibration created different groups of G and K stars within the HR diagram. A comparison between various photometrically calibrated luminosities and Hipparcos luminosities was presented by Jordi et al. (1997) for FGK dwarfs and subdwarfs. A preliminary new calibration was presented, removing the metallicity related drifts from existing calibrations. The remaining rms of the residuals between 388 F. VAN LEEUWEN the new photometrically calibrated luminosities and those obtained with Hipparcos is of the order of 0.3 mag, a factor two reduction with respect to pre-Hipparcos calibrations (see Figure 123). A similar study for B-type stars, using the Geneva photometric system, has been presented by Cramer (1997). All of these detailed luminosity calibrations provided by Hipparcos are expected to contribute significantly to the study of stellar structure and evolution, providing a much more detailed observational grid to compare theoretical isochrones with. 12.5. VARIABLE STARS The greatest impact of the Hipparcos results for variable stars is probably among the eclipsing binaries, where the known number of such stars as contained in the Hipparcos Catalogue was almost doubled (from 574 to 917), and among short period and small-amplitude variables of various types. The statistical aspects of the Hipparcos variable star results will be of great importance, in particular in better defining the regions of instability in the HR-diagram (Eyer & Grenon, 1997). An examination of Figure 93 shows clearly the main contribution of Hipparcos for periodic stars as a function of the photometric periods: between 0.1 and 8 days, with a peak around 1 day. This peak is not only the result of the difficulty for ground-based observations to detect regular variability on a time scale of 1 day, but also of observing strategy used by Hipparcos, which made detection of periods of about 1 day relatively easy. An example of a major contributor to this group can be seen in Figures 121 and 124, showing the Hipparcos amplitudes and periods for low amplitude periodic variables among A and F type stars. Although classifications were not clear at the time of the preparation for the Hipparcos Catalogue (and given the very large number of variable stars that had to be examined, there was not really time for detailed classifications) it is now clear that the concentration of stars with periods less than 0.2 d and amplitudes below 0.1 mag are probably all Scuti stars, while those with periods longer than 0.4 d and amplitudes less than 0.1 mag are most likely variable Ap stars and related types. Noteworthy seems to be a group of spectral type F0 stars with periods between 0.6 and 2 days, many of which are of luminosity class III or IV. Among the A and F stars in these groups are likely to be a small number of low amplitude close binary systems, which have not yet been properly recognized. An example is the early A-type star HIP 85501, which could, judging on the shape of the minimum, be an eclipsing binary with a 2.215 d period. It is clear that there is still a lot of research to be done on these stars to explain their various types of variability, and the way this may be connected to their structural parameters and evolutionary status. Various studies on semi-regular, small- and large-amplitude variables have started already (for Cyg stars by van Leeuwen et al., 1997d, van Leeuwen et al., 1997d, for B and Be stars by Hubert et al., 1997, for Herbig Ae/Be stars by van den Ancker et al., 1997). A possible detection of rapid variations in Mira stars by de Laverny 389 THE HIPPARCOS MISSION Figure 124. Periods, types and amplitudes for low amplitude intrinsic variables of spectral types A and F, many of which were discovered by Hipparcos. The meaning of the symbols is as follows (bearing in mind however the uncertainties in Hipparcos classifications): open circles: Scuti; filled circles: SX Phe; open stars: RRc; filled stars: RRab; open triangles: Ap stars with no variability type; filled triangles: remaining A stars; : remaining F stars. + 8000 8500 9000 BJD - 2 440 000 Figure 125. The Hipparcos observations of the Mira T Cas (HIP 1834), shown with respect to the fitted and transformed AAVSO observations (dotted line), in the way these stars are presented in Volume 12 of the Hipparcos Catalogue. et al. (1997) was deduced from the Hipparcos data, but has to be treated with some care, in particular as these variations take place near minimum light, which makes the Hipparcos photometric data much more vulnerable to incidental superpositions. This can show up in particular through a large difference between the magnitudes as measured in the preceding and following fields of view. The scanning mode and observing frequency for Hipparcos was not well suited for the slowly varying red variables (Miras, semi-regulars and slow irregulars). This was compensated by a collaboration between the Hipparcos project and the 390 F. VAN LEEUWEN Figure 126. Period versus amplitude for M-Miras, showing the upper limit of the amplitudes as a function of period. (from Mattei et al., 1997) AAVSO, which led to an unique data base for astrometric and photometric information on these stars (Mattei et al., 1997, Mennessier et al., 1997). During the Hipparcos mission, this data base was used in planning the observations of slowly varying large-amplitude variables (Mennessier et al., 1992). The combination of the AAVSO and Hipparcos observations is shown in Volume 12 of ESA (1997), in Section B of the folded light curves. There, the Hipparcos data are shown with respect to the fitted light curves of the AAVSO data, carefully transformed to the same system by M. Grenon. An example is shown in Figure 125. The AAVSO data base contains extensive visual data for 255 Mira and semiregular variables that were observed by the Hipparcos satellite, with for most at least 1000 observations, and for many in excess of 10 000 observation, covering a time span from October 1961 till August 1996. Most of these stars, 170, are classified as Mira-type, the remainder are types of SR variables. Various relations between periods, amplitudes and variation of amplitudes were used by Mattei et al. (1997) to re-examine the classification of these stars, leading to 7 revised classifications: 5 SR-types that are in fact Miras, and two Miras that are SRa-types. Also recognized in the data is an upper limit to the amplitudes of Miras as a function of their periods, expressed as: A 5:72 log P , 7:85 mag: (106) THE HIPPARCOS MISSION 391 Figure 127. Periods, types and amplitudes for eclipsing variables. Upper diagram: new discoveries; lower diagram: previously known variables. Open circles: EW types; filled circles: EB types; open stars: EA types; crosses: ELL types; filled stars: probably eclipsing, but insufficient coverage for type determination. This relation is shown in Figure 126. The analysis also revealed a group of 30 stars having two periods with a ratio between 1.7 and 1.95, below the expected ratio. Most of these are SRb types, and 3 are Miras. A different grouping of Miras was done by Mennessier et al. (1997). Using the “LM method”, the stars were divided into kinematical groups, for which the mean luminosity, distribution and velocity characteristics were calibrated. It was shown how the different types of red variables may be related to these kinematical groups, which indirectly relate types of long period red variables to stars of different age and mass. A very rich source of material may become available through the discovery of over 300 eclipsing binaries. Most of the new discoveries are in the lower amplitude range and for periods between 0.8 and 3.0 d, a range which is difficult to cover with ground-based measurements. The first paper on one of the newly discovered EB type binaries ( CMa) was presented by van Leeuwen & van Genderen (1997). The notes accompanying the variable star annexes in the Hipparcos Catalogue show 392 F. VAN LEEUWEN many cases where existing ephemerides have been compared with the timing of minimum light as measured by Hipparcos, in most cases leading to a more accurate period determination. In one case at least, notably for Lyr, the Hipparcos light curve confirmed the changing period as given by Sahade (1980). 12.6. REFERENCE FRAME IMPROVEMENTS The first proof of the impact of the improvement in the accuracy of the stellar reference frame was obtained through a comparison between radio and optical images for SN1987A (Reynolds et al., 1995). The alignment of the Hubble Space Telescope image with the radio map to an accuracy of at least 0.1 arcsec made it possible to understand much better and with greater confidence the relation between the radio and optical signals. An extreme example of the effect of a positional improvement was presented by Caraveo et al. (1997) for the mV = 25:5 optical counterpart of the Geminga radioquiet pulsar. Connecting through 5 steps the Hipparcos and Tycho positions to the HST image of the pulsar, a positional accuracy of 0.04 arcsec was obtained. This accuracy was required to properly relate the observations to the barycentre of the solar system, and thus combine the observations from SAS-2, COS-B and EGRET, so that the second time derivative in the pulsar frequency could be determined. The procedures described above rely on the possibility to obtain much improved absolute positions from photographic plates, using the Hipparcos and Tycho reference frames (Robichon et al., 1995, Platais et al., 1995). It was shown for plates taken at an epoch close to the Hipparcos reference epoch of J1991.25, that the rms of the residuals could be significantly reduced with respect to pre-Hipparcos reference frames (from 0.3 arcsec down to 0.07 arcsec accuracy for positions on a Schmidt plate). High order polynomials were, however, needed to represent the distortions of positions on a Schmidt plate. The possibilities for smaller astrograph plates are less favourable due to the smaller numbers of reference stars available. Some of these problems can be overcome when many plates of different parts of the sky are available, and an overall field distortion pattern can be determined (Zacharias et al., 1997). Using the Hipparcos and Tycho Catalogues for a new reduction of the photographic plates used for the CPC2 catalogue, the accuracies of the derived positional coordinates was in all cases significantly improved. The same conclusion was reached from the implementation of the Hipparcos positions as reference points in meridian circle observations by Réquième et al. (1997). In comparison with using PPM reference stars, the mean standard errors dropped from 0.3 arcsec to 0.06 arcsec, with accuracies for the measured positions better than 0.1 arcsec for stars brighter than mV = 14:5. The much improved star positions and magnitudes are also already assisting satellites currently in operation, such as ISO, to improve the mission operations 393 THE HIPPARCOS MISSION Table XXVII DE200: FAST solution for the rotation and spin parameters. Values of are given in mas, and values of ! in mas/yr. 1 2 3 !1 !2 !3 ,11.90 ,12.64 4.32 ,9.51 14.91 2.62 4.12 1.18 1.47 3.68 3.71 2.00 (Batten et al., 1997) and are already incorporated in the planning of future satellites such as ENVISAT (Ratier et al., 1997) and AXAF (Green et al., 1997). 12.7. SOLAR SYSTEM RELATED STUDIES The duration of the Hipparcos mission was rather short in comparison with orbital periods in the solar system. The impact of the solar system observations will therefore be rather limited, providing in most cases a few high accuracy reference points. However, improvements for orbital parameters have been obtained by Viateau & Rapaport (1997) through incorporating Hipparcos data in ground-based data for two asteroids, Pallas and Bamberga. The improvements were most significant for the inclinations and nodes of the orbits. A determination of the mass of Ceres, using disturbances to these orbits and incorporating Hipparcos data for improved orbit descriptions, showed a mass some 5 per cent lower than the value adopted by the IAU. The new value for the mass is: MCeres = (4:785 0 039)10,10M : : (107) In a different approach, examining the instantaneous effects of close encounters between asteroids, Bange & Bec-Borsenberger (1997) derived masses for a small number of asteroids, though with relatively low accuracy. In the same paper, the link between the dynamical reference frame given by the DE200 ephemerides and the ICRS was determined, using the minor-planet data, to an accuracy of a few mas. The rotation and spin values are given in Table XXVII. Further details on the methods used for solving these parameters are given by Bec-Borsenberger et al. (1995) and Bougeard et al. (1997). Indirect improvements of information on orbits in the solar system can be obtained through the use of the Hipparcos reference positions in the reduction of existing plate material (Fienga et al., 1997). Improvements by a factor 5 to 10 were obtained from a number of experiments, using only small numbers of Hipparcos stars per plate. An alternative way of interpreting the Hipparcos data on solar system objects was presented by van Leeuwen & Evans (1997) and Morrison et al. (1997b). In an application to the ephemerides of Jupiter and Saturn (measured through the positions of the moons Europa, Titan and Callisto), a small number of reference 394 F. VAN LEEUWEN Figure 128. Comparison of the observation positions of Jupiter with DE200 and DExxx ephemerides in right ascension (upper graph) and declination (lower graph). The Hipparcos normal points derived from observations of Europa are shown as filled squares. Individual Tycho observations derived from observations of Calisto are plotted as light error bars. The opposition means derived from Carlsberg Meridian Telescope observations are plotted as heavy error bars. (from Morrison et al., 1997) points was constructed from the available data. These reference points are twodimensional rather than the one-dimensional observations, and can be used as (very accurate) positional determinations in combination with ground-based measurements in new ephemerides calculations. The data confirmed the discrepancies observed with the Carlsberg Meridian telescope between the DE200 ephemerides and the measured positions. An example of these ephemerides comparisons, including a preliminary improvement of DE200 (DExxx), is shown in Figure 128. 13. Future projects The scientific basis for further astrometric missions, obtaining higher accuracies and measuring fainter stars, is very strong. A wide range of fundamental topics in astronomy can be covered, ranging from stellar structure and stellar evolution through kinematics, masses and cluster dynamics to understanding of our Galaxy, extragalactic astrophysics and general relativity, while planetary searches, supernovae detections and a host of other scientific objectives become feasible too (Perryman et al., 1997c, Peterson & Shao, 1997). The almost automatic “byproduct” of a photometric survey adds to this the discovery and surveying of any detectable variable stars, among which there will be thousands of Cepheid, RR Lyrae, Mira and other pulsating stars, for all of which very well determined distances can be obtained. The most substantial new astrometric survey is planned by the GAIA mission, for which ESA has recently started the first studies. GAIA THE HIPPARCOS MISSION 395 aims at obtaining astrometric parameters to 1 to 10 arcsec accuracy for more than 100 million objects, and can be complemented with a survey down to 18-20 mag, with astrometric parameters to accuracies 10 times better than Hipparcos. Hipparcos has led the way and has shown that obtaining high quality astrometric data from space is not only possible, but reaches in general far higher accuracies. It is also much more efficient than obtaining similar data from the ground. Hipparcos has, however, also shown some weaknesses: the relatively short length of the mission introduced an unwanted correlation between orbital motions and proper motions, an effect that will only get worse when accuracies get higher. More accurate missions should therefore be planned to last longer. The first research on the Hipparcos results exposed the need for supporting data, primarily detailed photometric measurements, radial velocities and spectral classifications. This was recognized well before the launch of the mission, but only limited support was received (from the ESO key programmes) for a large-scale radial velocity survey. Again, obtaining much larger amounts of data than Hipparcos, such deficiencies become only more severe, as much more automated processing and interpretation of the data is required. Improved photometric observations can be incorporated in a future mission (Straižys & Høg, 1995, Straižys et al., 1997), and there may be possibilities for spectral classifications and radial velocities too (Favata & Perryman, 1995, Favata & Perryman, 1997, Bastian, 1995, LaSala & Kurtz, 1995). Although there are some small advantages in a pointing satellite, in particular through the accumulation of observing time for specific objects, the scanning satellite, using two or more apertures, remains by far the most effective way of accumulating absolute parallaxes for large numbers of stars. Significant improvements in positional accuracy can be obtained incorporating interferometric techniques, such as planned for the GAIA design (Lindegren & Perryman, 1995) and in DIVA (Röser et al., 1997), a smaller, German version of the GAIA proposal, or LIGHT, a Japanese version of GAIA (Yoshizawa et al., 1997). A summary of the kind of problems faced in designing an astrometric interferometry space mission were outlined by Gai et al. (1997): detector efficiency, resolution and dynamic range, optical design, stability of the system, attitude control, are only a few of the problems that have to be considered from the point of view of the much higher requirements set by such missions in comparison with the Hipparcos satellite. As an example of the complexity of the data analysis, the investigations on variable stars are considered here. Given the very large numbers of stars to be observed by GAIA, and the two or possibly three interferometers, an estimated 1011 observations would be obtained for 108 stars, in 7 photometric channels, accompanied by parallax information for most of these objects, determining distances to a very high accuracy. Information on colours and distances (absolute magnitudes) can assist in variability classifications, which will have to be made unsupervised by computer software, possibly incorporating Bayesian methods. Due to the scanning law, however, the window functions for the period searches on those data remain very bad, in particular for time scales of 5 to 100 d. Small changes in the scanning 396 F. VAN LEEUWEN Figure 129. The coverage of the sky by Hipparcos over a period of 6 months. Two small areas near the ecliptic plane were not covered, which is reflected in the occurrences of gaps in the photometric data as shown in Figure 16. law can make large differences on the coverage. In particular, a larger angle in the scanning law will lead to a decrease in the lengths of the largest gaps in the data for stars near the ecliptic plane, and at the same time will produce a better overall coverage of scanning directions, which is of benefit to the double star analysis. A closing gap (a small part of the sky centered on the ecliptic plane that was not covered at all during a half-year scan, see Figure 129) existed for Hipparcos and introduced unnecessarily very large gaps in the photometric data for a small number of stars near to the ecliptic (see also Figure 16). Such gaps, which do much harm to automated period searches, should be avoided for any further astrometry mission. Contrary to the astrometric data, which are all interconnected and can therefore only be obtained at the end of a mission, the photometric data stand on their own. It would be advisable to consider early publication of details on variable stars discovered during a new mission, so that ground-based observations of the kind not provided by a mission can be obtained simultaneously with the further mission observations. Finally, with several missions being planned for more or less the same epoch, the combined interpretation of data should be considered, in particular for the astrometric data. Ultimately, one should be able to combine abscissa residuals from Hipparcos and future missions to obtain the best possible astrometric parameters. This applies in particular to proper motions and their possible disturbance by orbital motions. Thus, the publication of abscissa residuals as was done for Hipparcos should be considered very seriously for any further astrometric mission too. THE HIPPARCOS MISSION 397 Acknowledgements A large number of people worked together to make Hipparcos a success, and this paper tried to do some justice to their work. Above all should be mentioned the project scientist Michael Perryman and the consortia leaders Jean Kovalevsky, Erik Høg and Lennart Lindegren, and the remaining members of the Hipparcos science team as well as the various working groups. The team at ESOC, led by Dietmar Heger and Alastair McDonald, rescued the mission from what appeared to be an early exit. The three data reduction teams then succeeded to turn the mission into a success even well beyond what was expected for the nominal mission, not least thanks to an element of competition and the regular comparison exercises. Much use was made of the more detailed descriptions of the data presentation and reduction procedures in Volumes 1 to 4 of the Hipparcos Catalogue, of which Volume 1 contains a list of the many people that contributed to the mission from within and outside the Input Catalogue and Data Reduction Consortia. All of these people helped to make the Hipparcos mission a total success. Finally I like to thank Andrew Murray who introduced me into project 16 years ago, the members past and present of the NDAC consortium, and my wife Małgorzata, who at many occasions contributed, behind the scenes, to the success of the NDAC team at the RGO, and Michael Perryman, Hans Schrijver, Dafydd Evans, Margaret Penston, Rudolf Le Poole and Erik Høg who read (parts of) earlier versions of the text and provided numerous corrections. References van Altena, W.F., Lee, J.T., Hoffleit, E.D.: 1995, ‘General Catalogue of trigonometric stellar parallaxes, Fourth edition’, Yale University Observatory Alvarez, R., Mennessier, M.-O., Barthès, D., Luri, X., Mattei, J.A.: 1997, ‘Oxygen-rich Miras: nearinfrared luminosity calibrations, populations and period-luminosity relations’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 383-386 van den Ancker, de Winter, D., Tjin A Djie, H.R.E.: 1997, ‘Hipparcos photometry of Herbig Ae/Be stars’, in A&A, submitted Arenou, F., Lindegren, L., Frœschlé, M., Gómez, A.E., Turon, C., Perryman, M.A.C., Wielen, R.: 1995, ‘Zero-point and external errors of the Hipparcos parallaxes’, in A&A, 304, pp. 52-60 Arenou, F., Gómez, A.E.: 1997, ‘Unbiased luminosity calibrations for Hipparcos data’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 287-289 Bange, J.-F., Bec-Borsenberger, A.: 1997, ‘Determination of the masses of minor planets’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 169-172 Barraclough, D.R.: 1985, ‘International Geomagnetic Reference Field revision 1985’, in Pure Appl.Geophysics, 123, pp. 641-645 Bässgen, G., Wicenec, A., Andreasen, G.K., Høg, E., Wagner, K., Wesselius, P.: 1992, ‘Tycho transit detection’, in A&A, 258, pp. 186-192 Bastian, U.: 1995, ‘100 million radial velocities: how to get them’, in Future possibilities for astrometry in space, M.A.C. Perryman, F. van Leeuwen (eds.), ESA, SP-379, pp. 165-167 Bastian, U., Born, E., Agerer, F., Dahm, M., Großmann, V., Makarov, V.: 1996, ‘Confirmation of the classification of a new Tycho variable: HD 32456 is a 3.3-day Cepheid’, in IBVS, 4306 398 F. VAN LEEUWEN Batten, A., Marc, X., McDonald, A., Schütz, A.: 1997, ‘The use of the Hipparcos and Tycho catalogues in flight dynamics operations at ESOC’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 191-194 Bec-Borsenberger, A., Bange, J.-F., Bougeard, M.-L.: 1995, ‘Hipparcos minor planets: First step towards the link between the Hipparcos and the dynamical reference frames’, in A&A, 304, pp. 176-181 Benedict, G.F., Nelan, E., Story, D., McArthur, B., Whipple, A.L., Jefferys, W.H., van Altena, W., Hemenway, P.D., Shelus, P.J., Franz. O.G., Bradley, A., Frederick, L.W., Duncombe, R.L.: 1992, ‘Astrometric performance characteristics of the Hubble Space Telescope fine guidance sensors’, in PASP, 104, pp. 958-975 Bernstein, H.-H.: 1997, ‘Astrometric indications of brown dwarfs based on Hipparcos data’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 705-708 Bernstein, H.-H., Bastian, U.: 1995, ‘Finding planets and brown dwarfs with GAIA’, in Future possibilities for astrometry in space, M.A.C. Perryman, F. van Leeuwen (eds.), ESA, SP-379, pp. 55-59 Bessel, F.W.: 1841, ‘Beobachtungen verschiedener Sterne der Plejaden’ in Astronomische Untersuchungen, Erster Band, pp. 209-238 Le Betre, T., Epchtein, N., de Batz, B., Bertin, E., Borsenberger, J., Copet, E., Deul, E., Forveille, T., Fouqué, P., Guglielmo, F., Holl, A., Hron, J., Kienel, C., Kimeswenger, S., Lacombe, F., Le Sidaner, P., Rouan, D., Ruphy, S., Schultheis, M., Simon, G., Tiphène, D.: 1995, ‘The Deep near-infrared survey of the southern sky (DENIS)’, in ESO/ST-ECF workshop on calibrating and understanding HST and ESO instrumentation, pp. 195-204 Bierman, G.J.: 1977, ‘Factorization Methods for discrete sequential estimation’, Mathematics in Science and Engineering, Vol. 128, N.Y.: Academic Press Binney, J.J., Tremaine, S.: 1987, Galactic Dynamics, Princeton Univ.Press, Princeton Binney, J.J., Dehnen, W., Houk, N., Murray, C.A., Penston, M.J.: 1997, ‘The kinematics of mainsequence stars from Hipparcos data’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 473-477 de Boer, K.S., Schmidt, J.H.K., Heber, U.: 1995, ‘Hot HB stars in globular clusters: physical parameters and consequences for theory. II. NGC 6397 and its short blue horizontal branch’, in A&A, 303, pp. 95-106 de Boer, K.S., Tucholke, H.-J., Schmidt, J.H.K.: 1997a, ‘Calibrating horizontal-branch stars with Hipparcos’, in A&A, 317, pp. L23-L26 de Boer, K.S., Geffert, M., Tucholke, H.-J., Schmidt, J.H.K.: 1997b, ‘Calculating the mass of horizontal-branch stars with Hipparcos’, in Hipparcos, Venice’97, M.A.C. Perryman, P.L. Bernacca (eds.), ESA, SP-402, pp. 331-334 Bois, E.: 1986, ‘First-order theory of stellar attitude motion. Application to Hipparcos’, in Cel.Mech., 39, pp. 309-327 Borriello, L.: 1987, ‘Imaging approach to multiple star recognition’, in The third FAST thinkshop, P.L. Bernacca, J. Kovalevsky (eds.), Padova, pp. 189-197 Borriello, L., Delaney, W.: 1985, ‘Double star detection’, in The second FAST thinkshop, J. Kovalevsky (ed.), Marseille, pp. 159-167 Bougeard, M.L., Bange, J.-F., Caquineau, C., Bec-Borsenberger, A.: 1997, ‘Robust estimation with application to the Hipparcos minor planet data’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 165-168 Brosche, P., Geffert, M.: 1988, ‘A small extragalactic link for Hipparcos’, in Scientific aspects of the Input Catalogue preparation II, J. Torra, C. Turon (eds.), pp. 475-479 Brown, A.G.A., Arenou, F., van Leeuwen, F., Lindegren, L., Luri, X.: 1997a, ‘Some considerations in making full use of the Hipparcos catalogue’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 63-68 Brown, A.G.A., Perryman, M.A.C., Kovalevsky, J., Robichon, N., Turon, C., Mermilliod, J.-C.: 1997b, ‘The Hyades: distance, structure and dynamics’, in Hipparcos, Venice’97, M.A.C. Perryman, P.L. Bernacca (eds.), ESA, SP-402, pp. 681-686 Brown, P.J.F., Dufton, P.L., Lennon, D.J., Keenan, F.P.: 1986, ‘The chemical composition of six southern clusters and associations’, in MNRAS, 220, pp. 1003-1020 THE HIPPARCOS MISSION 399 de Bruijne, J.H.J., Hoogerwerf, R., Brown, A.G.A., Aguilar, L.A., de Zeeuw, P.T.: 1997, ‘Improved methods for identifying moving groups’, in Hipparcos, Venice’97, M.A.C. Perryman, P.L. Bernacca (eds.), ESA, SP-402, pp. 575-578 Burnage, R., Fehrenbach, Ch., Duflot, M.: 1988, ‘Vitesse radiale pour Hipparcos’ in Scientific aspects of the Input Catalogue preparation II, J. Torra, C. Turon (eds.), pp. 427-431 Caldwell, J.A.R., Laney, C.D.: 1991, ‘Cepheids in the Magellanic Clouds’, in The Magellanic Clouds, R.Haynes, D.Milne (eds.), IAU Symp. 148, Dordrecht: Kluwer Academic Publishers, pp. 259-257 Caraveo, P.A., Lattanzi, M.G., Massone, G., Mignani, R., Makarov, V.V., Perryman, M.A.C., Bignami, G.F.: 1997, ‘The importance of positioning isolated neutron stars: the case of Geminga’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 709-714 Cayrel de Strobel, G., Crifo, F., Lebreton, Y.: 1997, ‘The impact of Hipparcos on the old problem of the Helium content of the Hyades’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 687-688 Connes, P.: 1979, ‘Should we go to space for parallaxes?’, in A&A, 71, pp. L1-L4 Cramér, H.: 1946, ‘Mathematical methods of statistics’, Princeton University Press Cramer, N.: 1997, ‘Absolute magnitude calibrations of Geneva photometry based on Hipparcos parallaxes: the B-type stars’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 311-313 Crézé, M., Chereul, E., Bienaymé, O., Pichon, C.: 1997, ‘The distribution of nearby stars in phase space mapped by Hipparcos: the potential well and local dynamical mass’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 669-674 Daly, E.J., van Leeuwen, F., Evans, H.D.R, Perryman, M.A.C.: 1994, ‘Radiation-belt and transient solar-magnetospheric effects on Hipparcos radiation background’, in IEEE Trans.nucl.science, Vol. 41, No. 6, Part 1, pp. 2376-2382 Deeming, T.J.: 1975, ‘Fourier analysis with unequally-spaced data’, in Astroph.&Space Sci., 36, pp. 137-158 Delaney, W.: 1983, ‘The double star intensity model: formal solutions’, in The FAST thinkshop: Processing of the scientific data from the ESA astrometric satellite Hipparcos, P.L. Bernacca (ed.), Padova, pp. 217-219 Delaney, W.: 1985, ‘Double star analysis strategy’, in The second FAST thinkshop, J. Kovalevsky (ed.), Marseille, pp. 153-158 Delaney, W.: 1987, ‘Double star signal analysis and knowledge aquisition strategy’, in The third FAST thinkshop, P.L. Bernacca, J. Kovalevsky (eds.), Padova, pp. 179-188 Delhaye, J.: 1965, ‘Solar motion and velocity dispersion of common stars’, in Stars and stellar systems, V, A. Blaauw, M. Schmidt (eds.), The University of Chicago Press, Chicago, pp. 61-84 Dick, W.R., Ruben, G., Schilbach, E., Scholz, R.-D.: 1987, ‘A program to derive from terrestrial observations the rotation of the Hipparcos system with reference to the inertial system’, in Astron.Nachr., 308, pp. 211-216 Dommanget, J.: 1985, ‘Concerns, realisations and prospects of the sub-group double star of the Hipparcos INCA consortium, I.’, in Scientific aspects of the input catalogue preparation, ESA, SP-234, pp. 153-159 Dommanget, J.: 1988, ‘Concerns, realisations and prospects of the sub-group double star of the Hipparcos INCA consortium, II.’, in Scientific aspects of the input catalogue preparation II, J. Torra, C. Turon (eds.), pp. 191-206 Dommanget, J., Nys, O.: 1982, ‘Second catalogue of ephemerides’, in Observatoire Royal de Belgique, Communications, Serie B, No 124 Dommanget, J., Nys, O.: 1994, ‘Catalogue of the Components of Double and Multiple stars (CCDM)’, in Observatoire Royal de Belgique, Communications, Serie A, No 11, pp. 1-25 Donati, F., Frœschlé, M., Falin, J.-L., Canuto, E., Kovalevsky, J.: 1992, ‘Attitude determination in the Hipparcos revised mission’, in A&A, 258, pp. 41-45 Donati, F., Sechi, G.: 1992, ‘Method of comparison between determinations of the Hipparcos attitude’, in A&A, 258, pp. 46-52 Dravins, D., Lindegren, L., Madsen, S., Holmberg, J.: 1997, ‘Astrometric radial velocities from Hipparcos’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 733-738 400 F. VAN LEEUWEN Egret, D., Dideleon, P., McLean, B.J., Russell, J.L., Turon, C.: 1992, ‘The Tycho input catalogue. Cross-matching the guide star catalogue with the Hipparcos INCA data base’, in A&A, 258, pp. 217-222 ESA: 1989, ‘The Hipparcos mission, pre-launch status’, ESA, SP-1111 ESA: 1992, ‘The Hipparcos Input Catalogue’, ESA, SP-1136 ESA: 1997, ‘The Hipparcos and Tycho Catalogues’, ESA, SP-1200 Evans, D.W.: 1992, ‘The APM Proper Motion Project - I. High proper motion stars’, in MNRAS, 255, pp. 521-538 Evans, D.W., van Leeuwen, F., Penston, M.J., Ramamani, N., HøG, E.: 1992, ‘Hipparcos photometry: NDAC reductions’, in A&A, 258, pp. 149-156 Evans, D.W., Irwin, M.J.: 1995, ‘The APM Proper Motion Project - II. The lower proper motion stars’, in MNRAS, 277, pp. 820-844 Eyer, L., Grenon, M.: 1997, ‘Photometric variability in the HR-diagram’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 467-471 Farilla, A.: 1985, ‘Estimation of double star parameters’, in The second FAST thinkshop, J. Kovalevsky (ed.), Marseille, pp. 169-173 Favata, F., Perryman, M.A.C.: 1995, ‘Parallel acquisition of radial velocities and metallicities for a GAIA-type interferometric astrometry mission’, in Future possibilities for astrometry in space, M.A.C. Perryman, F. van Leeuwen (eds.), ESA, SP-379, pp. 153-163 Favata, F., Perryman, M.A.C.: 1997, ‘Parallel acquisition of radial velocities and metallicities for the GAIA mission’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 771-775 Feast, M.W., Catchpole, R.M.: 1997, ‘The Cepheid period-luminosity zero-point from Hipparcos trigonometric parallaxes’, in MNRAS, 286, pp. L1-L5 Fehrenbach, C., Duflot, M., Mannono, C., Burnage, R., Genty, V.: 1997, ‘Radial velocities. VIII. Ground-based measurements for Hipparcos’, in A&AS, 124, pp. 255-257 Fernley, J., Barnes, T.G., Skillen, I., Hawley, S.L., Hanley, C.J., Evans, D.W., Solano, E., Garrido, R.: 1997, ‘Absolute magnitudes of RR Lyrae stars’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 635-638 Ferraz-Mello, S.: 1981, ‘Estimation of period from unequally spaced observations’, in AJ, 86, pp. 619624 Fienga, A., Arlot, J.E., Pascu, D.: 1997, ‘Impact of Hipparcos data on astrometric reduction of solar system bodies’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 157-159 Florkowski, D.R., Johnston, K.J., Wade, C.M., de Vegt, C.: 1985, ‘Stellar radio astrometry. I. Precise positions of the radio emission of twenty stars’, in AJ, 90, pp. 2381-2385 Fricke, W., Schwan, H., Ledere, T.: 1989, ‘FK5’, Veröffentlichungen Astronomisches Rechen Institut Heidelberg Vol. 32 Gai, M., Bertinetto, F., Bisi, M., Canuto, E., Carollo, D., Cesare, S., Lattanzi, M.G., Mana, G., Thomas, E., Viard, T.: 1997, ‘GAIA feasibility: current research on critical aspects’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 835-838 Gazengel, F., Spagna, A., Pannunzio, R., Morbidelli, R., Sarasso, M.: 1995, ‘The Hipparcos multiple stars: reduction methods and preliminary results’, in A&A, 304, pp. 105-109 Geffert, M., Klemola, A.R., Hiesgen, M., Schmoll, J.: 1997, ‘Absolute proper motions for the calibration of the Hipparcos proper motion system’, in A&AS, 124, pp. 157-161 Gliese, W.: 1969, ‘Catalogue of nearby stars, edition 1969’, in Veröff.Astron.Rechen Inst. Heidelberg, 22, pp. 1-117 Gliese, W.: 1979, ‘Nearby star data published 1969-1978’, in A&AS, 38, pp. 423-448 Gliese, W., Upgren, A.R.: 1990, ‘Parallaxes: History, survey and outlook’, in Publ. Obs. Astron. Belgrade, 40, pp. 7-18; (also: Astronomisches Rechen-Institut Heidelberg, Mitteilungen Serie A, 225) Gómez, A.E., Luri, X., Mennessier, M.O., Torra, J., Figueras, F.: 1997a, ‘The luminosity calibration of the HR diagram revisited by Hipparcos’, in Hipparcos, Venice’97, M.A.C. Perryman, P.L. Bernacca (eds.), ESA, SP-402, pp. 207-212 THE HIPPARCOS MISSION 401 Gómez, A.E., Grenier, S., Udry, S., Haywood, M., Meillon, L., Sabas, V., Morin, D.: 1997b, ‘Kinematics of disk stars in the solar neighbourhood’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 621-623 Green, P.J., Aldcroft, T.A., Garcia, M.R., Slane, P., Vrtilek, J.: 1997, ‘Using the Tycho catalogue for AXAF: guiding and aspect reconstruction for half-arcsecond X-ray images’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 187-190 Greisen, E.W., Harten, R.H.: 1981, ‘An extension of FITS for groups of small arrays of data’, in A&AS, 44, pp. 371-374 Grenon, M.: 1984, ‘The physical properties of 1250 stars from the Gliese catalogue’, in The nearby stars and the stellar luminosity function, IAU coll.76, A.G. Davis Philip, A.R. Upgren (eds.), L.Davis Press, Schenectady, NY, pp. 439-446 Grenon, M.: 1988, ‘The new ground based photometric measurements’, in Scientific aspects of the input catalogue preparation II, J. Torra, C. Turon (eds.), pp. 343-354 Grenon, M., Mermilliod, M., Mermilliod, J,-C.: 1992, ‘The Hipparcos Input Catalogue. III. Photometry’, in A&A, 258, pp. 88-93 Großmann, V., Bässgen, G., Evans, D.W., Grenon, M., Grewing, M., Halbwachs, J.L., Høg, E., Mauder, H., Snijders, M.A.J., Wagner, K., Wicenec, A.: 1995, ‘Results on Tycho photometry’, in A&A, 304, pp. 110-115 Halbwachs, J.-L., Høg, E., Bastian, U., Hansen, P.C., Schwekendiek, P.: 1992, ‘Tycho star recognition’, in A&A, 258, pp. 193-200 Hansen-Ruiz, C.S., van Leeuwen, F.: 1997, ‘Definition of the Pleiades main sequence in the Hertzsprung-Russell diagram’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 295-297 Harris, H.C., Dahn, C.C., Monet, D.G.: 1997, ‘Accurate ground-based parallaxes to compare with Hipparcos’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 105-107 Hemenway, P.D., Benedict, G.F., Jefferys, W.H., Shelus, P.J., Duncombe, R.L.: 1988, ‘The extragalactic link. Operational preparations for the Hubble space telescope observations’, in Scientific aspects of the Input Catalogue preparation II, J. Torra, C. Turon (eds.), pp. 461-467 Hertzsprung, E.: 1907, ‘Zur Bestimmung der Photographischen Sterngröße’ in Astron.Nachr. Vol. 176, pp. 49-53 Hertzsprung, E.: 1952, ‘Accuracy of Yale parallaxes’, in The Observatory, 72, pp. 242-243 Hestroffer, D., Morando, B., Mignard, F., Bec-Borsenberger, A.: 1995, ‘Astrometry of minor planets with Hipparcos’, in A&A, 304, pp. 168-175 Høg, E.: 1960, ‘Proposal for a photoelectric meridian circle and a new principle of instrument design’, Astr.Abh.Sternw.Hamburg-Bergedorf, 5, pp. 264-272 Høg, E.: 1979, ‘The ESA astrometry satellite’, in Modern Astrometry, IAU Coll. 48, F.V. Prochazka, R.H. Tucker (eds.), University Observatory Vienna, pp. 557-559 Høg, E., Jaschek, C., Lindegren, L.: 1982, ‘TYCHO - A planned astrometric and photometric survey from space’, in: The scientific aspects of the Hipparcos space astrometry mission M.A.C.Perryman, T.D.Guyenne (eds.), ESA, SP-177, pp. 27-28 Høg, E., Bastian, U., Egret, D., Grewing, M., Halbwachs, J.L., Wicenec, A., Bässgen, G., Bernacca, P.L., Donati, F., Kovalevsky, J., van Leeuwen, F., Lindegren, L., Pedersen, H., Perryman, M.A.C., Petersen, C., Scales, D., Snijders, M.A.J., Wesselius, P.R.: 1992a, ‘Tycho data analysis. Overview of the adopted reduction software and the first results’, in A&A, 258, pp. 177-185 Høg, E., Bastian, U., Hansen, P.C., van Leeuwen, F., Lindegren, L., Pedersen, H., Saust, A.B., Schwekendiek, P., Wagner, K.: 1992b, ‘Tycho astrometry calibration’, in A&A, 258, pp. 201-205 Høg, E.: 1993, ‘Astrometry and photometry of 400 million stars brighter than 18 mag’, in Developments in astrometry and their impact on astrophysics and geodynamics, I.I. Mueller, B. Kołaczek (eds.), IAU Symp. 156, Dordrecht: Kluwer Academic Publishers, pp. 37-45 Høg, E., Lindegren, L.: 1994, ‘ROEMER satellite project: the first high-accuracy survey of faint stars’, in Galactic and solar system optical astrometry, L.V. Morrison, G.F. Gilmore (eds.), Cambridge University Press, pp. 246-252 402 F. VAN LEEUWEN Høg, E., Bastian, U., Halbwachs, J.L., van Leeuwen, F., Lindegren, L., Makarov, V.V., Pedersen, H., Petersen, C.S., Schwekendiek, P., Wagner, K., Wicenec, A.: 1995, ‘Tycho astrometry from half of the mission’, in A&A, 304, pp. 150-159 Høg, E., Bässgen, G., Bastian, U., Egret, D., Fabricius, C., Großmann, V., Halbwachs, J.-L., Makarov, V.V., Perryman, M.A.C., Schwekendiek, P., Wagner, K., Wicenec, A.: 1997, ‘The Tycho catalogue’, in A&A, 323, pp. L57-L60 Høg, E., Petersen, J.O.: 1997, ‘Hipparcos parallaxes and the nature of Scuti stars’, in A&A, 323, pp. 827-830 Hoogerwerf, R., de Bruijne, J.H.J., Brown, A.G.A., Lub, J., Blaauw, A., de Zeeuw, P.T.: 1997, ‘New results for three nearby OB associations’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 571-574 Houk, N., Cowley, A.P.: 1975, ‘Catalogue of two-dimensional spectral types for HD stars, Vol.1’, Department of astronomy, Univ.of Michigan, Ann Arbor Houk, N.: 1978, ‘Catalogue of two-dimensional spectral types for HD stars, Vol.2’, Department of astronomy, Univ.of Michigan, Ann Arbor Houk, N.: 1982, ‘Catalogue of two-dimensional spectral types for HD stars, Vol.3’, Department of astronomy, Univ.of Michigan, Ann Arbor Houk, N., Swift, C.M., Murray, C.A., Penston, M.J., Binney, J.J.: 1997, ‘The properties of mainsequence stars from Hipparcos data’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 279-282 Hubert, A.M., Floquet, M., Gómez, A.E., Aletti, V.: 1997, ‘Photometric variability of B and Be stars’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 315-318 Jahreiß, H., Réquième, Y., Argue, A.N., Dommanget, J., Rousseau, M., Lederle, T., Le Poole, R.S., Mazurier, J.M., Morrison, L.V., Nys, O., Penston, M.J., Périé, J.P., Prévot, L., Tucholke, H.J., de Vecht, C.: 1992, ‘The Hipparcos Input Catalogue. II. Astrometric Data’, in A&A, 258, pp. 82-87 Jahreiß, H., Wielen, R.: 1997, ‘The impact of Hipparcos on the catalogue of nearby stars: the stellar luminosity function and local kinematics’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 675-680 Jenkins, L.F.: 1952, ‘General Catalogue of Trigonometric Stellar Parallaxes’, Yale University Observatory. Jenkins, L.F.: 1963, ‘Supplement to the General Catalogue of Trigonometric Stellar Parallaxes’, Yale University Observatory. Johnston, K.J., Fey, A.L., Zacharias, N., Russell, J.L., Ma, C., de Vegt, C. et al.: 1995, ‘A radio reference frame’, in AJ, 110, pp. 880-915 Jordi, C., Masana, E., Luri, X., Torra, J., Figueras, F.: 1997, ‘Systematic effects on photometric parallaxes for FGK dwarfs and subdwarfs’, in Hipparcos, Venice’97, M.A.C. Perryman, P.L. Bernacca (eds.), ESA, SP-402, pp. 283-286 Kholopov, P.N. (ed.): 1985, ‘The general catalogue for variable stars’, Moscow, Nauka Publishing House Klemola, A.R., Hanson, R.B., Jones, B.F.: 1994, ‘Lick NPM program: NPM1 Catalog and its applications’, in Galactic and solar system optical astrometry, L.V. Morrison, G.F. Gilmore (eds.), Cambridge University Press, pp. 20-25 Kovalevsky, J.: 1984, ‘Prospects for space stellar astrometry’, in Space Sci.Rev, 39, pp. 1-63 Kovalevsky, J., Falin, J., Pieplu, J.L, Bernacca, P.L., Donati, F., Frœschlé, M., Galligani, I., Mignard, F., Morando, B., Perryman, M.A.C., Schrijver, H., van Daalen, D.T., van der Marel, H., Villenave, M., Walter, H.G. et al.: 1992, ‘The FAST Hipparcos Data Reduction Consortium: overview of the reduction software’, in A&A, 258, pp. 7-17 Kovalevsky, J., Lindegren, L., Frœschlé, M., van Leeuwen, F., Perryman, M.A.C., Falin, J.L., Mignard, F., Penston, M.J., Petersen, C.S., Bernacca, P.L., Bucciarelli, B., Donati, F., Hering, R., Høg, E., Lattanzi, M., van der Marel, H., Schrijver, H., Walter, H.G.: 1995, ‘Construction of the intermediate Hipparcos astrometric catalogue’, in A&A, 304, pp. 34-43 Kovalevsky, J.: 1996, ‘Concluding remarks’, in Astronomical and astrophysical objectives of submilliarcsecond optical astrometry, E. Høg, P.K. Seidelmann (eds.), IAU Symp. 166, Dordrecht: Kluwer Academic Publishers, pp. 409-432 THE HIPPARCOS MISSION 403 Kovalevsky, J., Lindegren, L., Perryman, M.A.C., Hemenway, P.D., Johnston, K.J., Kislyuk, V.S., Lestrade, J.F., Morrison, L.V., Platais, I., Röser, S., Schilbach, E., Tucholke, H.-J., de Vegt, C. et al.: 1997, ‘The Hipparcos Catalogue as a realisation of the extragalactic reference system’, in A&A, 323, pp. 620-633 Kron, R.G.: 1995, ‘The Sloan digital sky survey’, in Galaxies in the young universe, Lect.Notes Phys., 463, H. Hippelein, K. Meisenheimer, H.-J. Röser (eds.), pp. 260-264 Kroupa, P., Bastian, U.: 1997, ‘The Hipparcos proper motion of the Magellanic Clouds’, in New Astr., 2, pp. 77-90 Kuzmin, A., Bastian, U., Høg, E., Kuimov, K., Röser, S.: 1997, ‘Tycho reference catalogue: Pilot project results’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 125-128 Lacroute, P.: 1982, ‘Histoire du projet d’astrométrie spatiale’, in The scientific aspects of the Hipparcos space astrometry mission M.A.C.Perryman, T.D.Guyenne (eds.), ESA, SP-177, pp. 3-12 Lafter, J., Kinman, T.D.: 1965, ‘An RR Lyrae survey with the Lick 20-inch astrograph. II. The calculation of RR Lyrae periods by electronic computer’, in ApJS, 11, pp. 216-222 Lagrula, J.: 1903, ‘Étude sur les occultations d’amas d’étoiles par la lune avec un catalogue normal des Pléiades’ in Traveaux de l’observatoire de Lyon Vol. III, pp. 1-152 Lampens, P., Kovalevsky, J., Frœschlé, M., Ruymaekers, G.: 1997, ‘On the mass-luminosity relation’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 421-423 LaSala, J., Kurtz, M.J.: 1995, ‘Automated spectral classification and the GAIA project’, in Future possibilities for astrometry in space, M.A.C. Perryman, F. van Leeuwen (eds.), ESA, SP-379, pp. 169-172 de Laverny, P., Mennessier, M.O., Mignard, F., Mattei, J.A.: 1997, ‘Detection of rapid photometric variations in Mira-type variables’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 387-389 Lebreton, Y., Gómez, A.E., Mermilliod, J.-C., Perryman, M.A.C.: 1997, ‘The age and Helium content of the Hyades revisited’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 231-236 van Leeuwen, F.: 1985, ‘Proper motion studies of stars in and around open clusters’, in Dynamics of Star Clusters, J. Goodman, P. Hut (eds.), IAU Symp. 113, Dordrecht: Kluwer Academic Publishers, pp. 579-606 van Leeuwen, F.: 1995, ‘Sub-milliarcsecond astrometry of star clusters and associations’, in Astronomical and astrophysical objectives of sub-milliarcsecond optical astrometry, E. Høg, P.K. Seidelmann (eds.), IAU Symp. 166, Dordrecht: Kluwer Academic Publishers, pp. 227-232 van Leeuwen, F.: 1997a, ‘The Hipparcos photometry and variability annexes’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 19-24 van Leeuwen, F.: 1997b, ‘Application possibilities for the Hipparcos intermediate astrometric data’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 203-206 van Leeuwen, F., Evans, D.W., Lindegren, L., Penston, M.J., Ramamani, N.: 1992a, ‘Early improvements to the Hipparcos Input Catalogue through the accumulation of data from the satellite (including the NDAC attitude reconstruction description)’, in A&A, 258, pp. 119-124 van Leeuwen, F., Penston, M.J., Perryman, M.A.C., Evans, D.W., Ramamani, N.: 1992b, ‘Modelling the torques affecting the Hipparcos satellite’, in A&A, 258, pp. 53-59 van Leeuwen, F., Evans, D.W.: 1997, ‘On the use of the Hipparcos intermediate astrometric data’, A&AS in press van Leeuwen, F., Evans, D.W., Grenon, M., Großmann, V., Mignard, F., Perryman, M.A.C.: 1997a, ‘The Hipparcos mission: photometric data’, in A&A, 323, pp. L61-L64 van Leeuwen, F., Evans, D.W., van Leeuwen Toczko, M.B.: 1997b, ‘Statistical aspects of the Hipparcos photometric data’, in Statistical Challenges for Modern Astronomy, II, E. Feigelson, G.J. Babu (eds.), Springer Verlag, New York, in press. van Leeuwen, F., Feast, M.W., Whitelock, P.A., Yudin, B.: 1997c, ‘First results from Hipparcos trigonometric parallaxes of Mira-type variables’, in MNRAS, 387, pp. 955-960 van Leeuwen, F., van Genderen, A.M., Zegelaar, I.: 1997d, ‘Hipparcos photometry of 24 variable massive stars ( Cyg variables)’, in A&AS, in press 404 F. VAN LEEUWEN van Leeuwen, F., van Genderen, A.M.: 1997, ‘The discovery of a new massive O-type close binary: CMa (HD 57061), based on Hipparcos and Walraven photometry’, in A&A, in press van Leeuwen, F., Hansen-Ruiz, C.S.: 1997, ‘The parallax of the Pleiades cluster’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 689-692 Lestrade, J.-F., Phillips, R.B., Preston, R.A., Gabuzda, D.C.: 1992, ‘High-precision VLBI astrometry of th radio-emitting star CrB – a step in linking the Hipparcos and extragalactic reference frames’, in A&A, 258, pp. 112-115 Lestrade, J.-F., Jones, D.L., Preston, R.A., Phillips, R.B., Titus, M.A., Kovalevsky, J., Lindegren, L., Hering, R., Frœschlé, M., Falin, J.-L., Mignard, F., Jacobs, C.S., Sovers, O.J., Eubanks, M., Gabuzda, D.: 1995, ‘Preliminary link of the Hipparcos and VLBI reference frames’, in A&A, 304, pp. 182-188 Lindegren, L.: 1979a, ‘Detection and measurement of double stars with an astrometric satellite’, in Colloquium on satellite astrometry, C. Barbieri, P.L. Bernacca (eds.), Padova, pp. 117-124 Lindegren, L.: 1979b, ‘Photoelectric astrometry. A comparison of methods for precise image location’, in Modern Astrometry, IAU Coll. 48, F.V. Prochazka, R.H. Tucker (eds.), University Observatory Vienna, pp. 197-217 Lindegren, L.: 1980, ‘Atmospheric limitation to narrow-field optical astrometry’, in Mitt.Astr.Gesell., 48, p. 147 Lindegren, L., Söderhjelm, S.: 1985, ‘Sphere reconstitution by the Northern Data Analysis Consortium (NDAC)’, in The second FAST thinkshop, J. Kovalevsky (ed.), Marseille, pp. 237-242 Lindegren, L.: 1987a, ‘Hipparcos observations of minor planets and natural satellites: NDAC treatment’, in The third FAST thinkshop, P.L. Bernacca, J. Kovalevsky (eds.), Padova, pp. 285-293 Lindegren, L.: 1987b, ‘Treatment of double stars in the overall NDAC processing’, in The third FAST thinkshop, P.L. Bernacca, J. Kovalevsky (eds.), Padova, pp. 169-178 Lindegren, L.: 1988, ‘A correlation study of simulated Hipparcos astrometry’, in Scientific aspects of the Input Catalogue preparation II, J. Torra, C. Turon (eds.), pp. 179-188 Lindegren, L., van Leeuwen, F., Petersen, C.S., Perryman, M.A.C., Söderhjelm, S.: 1992a, ‘Positions and parallaxes from the Hipparcos satellite. A first attempt at a global astrometric solution’, in A&A, 258, pp. 134-141 Lindegren, L., Le Poole, R.S., Perryman, M.A.C, Petersen, C.S.: 1992b, ‘Geometrical stability and the evolution of the Hipparcos telescope’, in A&A, 258, pp. 35-40 Lindegren, L., Høg, E., van Leeuwen, F., Murray, C.A., Evans, D.W., Penston, M.J., Perryman, M.A.C., Petersen, C., Ramamani, N., Snijders, M.A.J., Söderhjelm, S., et al.: 1992c, ‘The NDAC Hipparcos data analysis consortium: Overview of the reduction methods’, in A&A, 258, pp. 18-30 Lindegren, L.: 1995, ‘Estimating the external accuracy of Hipparcos parallaxes by deconvolution’, in A&A, 304, pp. 61-68 Lindegren, L., Kovalevsky, J.: 1995, ‘Linking the Hipparcos Catalogue to the extragalactic reference system’, in A&A, 304, pp. 189-201 Lindegren, L., Perryman, M.A.C.: 1995, ‘The GAIA concept’, in Future possibilities for astrometry in space, M.A.C. Perryman, F. van Leeuwen (eds.), ESA, SP-379, pp. 23-32 Lindegren, L., Mignard, M., Söderhjelm, S., Badiali, M., Bernstein, H.-H., Lampens, P., Pannunzio, R., Arenou, F., Bernacca, P.L., Falin, J.-L., Frœschlé, M., Kovalevsky, J., Martin, C., Perryman, M.A.C., Wielen, R.: 1997, ‘Double star data in the Hipparcos catalogue’, in A&A, 323, pp. L53L56 Liu, Y.Y., Baglin, A., Auvergne, M., Goupil, M.J., Michel, E.: 1997, ‘The lower part of the instability strip: evolutionary status of delta Scuti stars’, in Hipparcos, Venice’97, M.A.C. Perryman, P.L. Bernacca (eds.), ESA, SP-402, pp. 363-366 Lub, J., Pel, J.W.: 1977, ‘Properties of the Walraven V BLUW photometric system’, in A&A, 54, pp. 137-158 Luri, X., Mennessier, M.-O., Torra, J., Figueras, F.: 1996, ‘A new maximum likelihood method for luminosity class determinations’, in A&AS, 117, pp. 405-415 Luri, X., Arenou, F.: 1997, ‘Utilisation of Hipparcos data for distance determinations: error, bias and estimation’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 449-452 THE HIPPARCOS MISSION 405 Lutz, T.E., Kelker, D.H.: 1973, ‘On the use of trigonometric parallaxes for the calibration of luminosity systems: theory’, in PASP, 83, pp. 573-578 Ma, C., Shaffer, D.B., de Vegt, C., Johnston, K.J., Russell, J.L.: 1990, ‘A radio-optical reference frame. I. Precise radio source positions determined by Mark III VLBI: Observations from 1979 to 1988 and a tie to the FK5’, in AJ, 99, pp. 1284-1298 Malmquist, K.G.: 1922, ‘On some relations in stellar physics’, in Meddel. Lund Obs., 100, pp. 1-52 Malmquist, K.G.: 1936, ‘The effect of an absorption of light in space upon some relations in stellar statistics’, in Meddel. Stockholm Obs., 26, pp. 1-9 van der Marel, H.: 1988, ‘On the “Great Circle Reduction” in the data analysis for the astrometric satellite Hipparcos’, PhD Thesis, Technical University Delft van der Marel, H., Petersen, C.S.: 1992, ‘Hipparcos great-circle reduction. Theory, results and intercomparisons’, in A&A, 258, pp. 60-69 Martin, C., Mignard, F.: 1997, ‘Masses of astrometric binaries’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 417-420 Mattei, J.A.: 1988, ‘The AAVSO and its variable star data bank’, in Scientific aspects of the input catalogue preparation II, J. Torra, C. Turon (eds.), pp.379-394 Mattei, J.A., Foster, G., Hurwitz, L.A., Malatesta, K.H., Willson, L.A., Mennessier, M.-O.: 1997, ‘Classification of red variables’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 269-274 Mennessier, M.O., Baglin, A.: 1988, ‘The variable stars in the Hipparcos mission’, in Scientific aspects of the input catalogue preparation II, J. Torra, C. Turon (eds.), pp. 361-378 Mennessier, M.-O., Barthès, D., Boughaleb, H., Figueras, F., Mattei, J.A.: 1992, ‘Computation of ephemerides for long-period variables for the Hipparcos mission’, in A&A, 258, pp. 99-103 Mennessier, M.-O., Mattei, J.A., Luri, X.: 1997, ‘New aspects of long-period variable stars from Hipparcos: first results’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 275-277 Mermilliod, J.-C., Turon, C., Robichon, N., Arenou, F., Lebreton, Y.: 1997, ‘The distance of the Pleiades and nearby clusters’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 643-650 Mignard, F., Frœschlé, M., Badiali, M., Cardini, D., Emanuele, A., Falin, J.-L., Kovalevsky, J.: 1992a, ‘Hipparcos double star recognition and processing within the FAST consortium’, in A&A, 258, pp. 165-172 Mignard, F., Frœschlé, M., Falin, J.-L.: 1992b, ‘Hipparcos photometry: FAST main mission reductions’, in A&A, 258, pp. 142-148 Mignard, F., Söderhjelm, S., Bernstein, H., Pannunzio, R., Kovalevsky, J., Frœschlé, M., Falin, J.-L., Lindegren, L., Martin, C., Badiali, M., Cardini, D., Emanuele, A., Spagna, A., Bernacca, P.L., Borriello, L., Prezioso, G.:, 1995, ‘Astrometry of double stars with Hipparcos’, in A&A, 304, pp. 94-104 Monet, D.G., Dahn, C.C., Vrba, F.J., Harris, H.C., Pier, J.R., Luginbuhl, C.B., Ables, H.D.: 1992, ‘U.S.Naval Observatory CCD parallaxes of faint stars, I. Program description and first results.’ in AJ, 103, pp. 638-665 Morrison, L.V., Argyle, R.W., Réquième, Y., Helmer, L., Fabricius, C., Einicke, O.H., Buontempo, M.E., Muiños, J.L., Rapaport, M.: 1990, ‘Comparison of FK5 with Bordeaux and Carlsberg meridian circle observations’ A&A, 240, pp. 173-177 Morrison, L.V., Garrington, S.T., Argyle, R.W., Davis, R.J.: 1997a, ‘Future control of the Hipparcos frame using MERLIN’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 143-145 Morrison, L.V., Hestroffer, D., Taylor, D.B., van Leeuwen, F.: 1997b, ‘Check on JPL DExxx using Hipparcos and Tycho observations’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 149-152 Murray, C.A., Argyle, R.W., Corben, P.M.: 1986, ‘A survey of trigonometric parallaxes and proper motions with the UK Schmidt telescope. II. Astrometric and photometric data for a complete sample of 6125 stars brighter than B=17.5, V=17.0 in the South Galactic Cap’, in MNRAS, 223, pp. 629-648 406 F. VAN LEEUWEN Murray, C.A.: 1986, ‘A survey of trigonometric parallaxes and proper motions with the UK Schmidt telescope. III. Stellar populations, density and luminosity distributions in the South Galactic Cap’, in MNRAS, 223, pp. 649-671 Murray, C.A., Penston, M.J., Binney, J.J., Houk, N.: 1997, ‘The luminosity function of main-sequence stars within 80 parsecs, from Hipparcos data’, in Hipparcos, Venice’97, M.A.C. Perryman, P.L. Bernacca (eds.), ESA, SP-402, pp. 485-488 Nemec, J.M., Lutz, T.E.: 1993, ‘Period-luminosity-metal abundance relations for Population II variable stars’, in New perspectives on stellar pulsation and pulsating variable stars, IAU coll. 139, J.M. Nemec, J.M. Matthews (eds.), Cambridge University Press, pp. 30-52 Neto, R.B.: 1989, ‘Bad day for astronomers’, in Nature, Vol.340, p. 491 North, P., Jaschek, C., Egret, D.: 1997, ‘Delta Scuti stars in the HR diagram’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 367-370 Pannunzio, R., Spagna, A., Lattanzi, M.G., Morbidelli, R., Sarasso, M.: 1992, ‘The treatment of Hipparcos observations of some peculiar double stars: anomalous cases’, in A&A, 258, pp. 173176 Papoulis, A.: 1991, ‘Probability, random variables and stochastic processes’, McGraw-Hill, New York Parenago, P.P.: 1950, ‘Investigations in the space velocities of stars’, in Astron.Z. USSR, 27, (in russian), pp. 150-165 Pel, J.W.: 1991, ‘A comparison of the systematic accuracies in four photometric systems’, in Precision photometry: Astrophysics of the galaxy, A.G.D. Davis Philip, A.R. Upgren, K.A. Janes (eds.), L.Davis Press, Schenectady, NY, pp. 165-172 Peterson, D., Shao, M.: 1997, ‘The scientific basis for the space interferometry mission’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 749-753 Perryman, M.A.C., Høg, E., Kovalevsky, J., Lindegren, L., Turon, C., Bernacca, P.L., Crézé, M., Donati, F., Grenon, M., Grewing, M., van Leeuwen, F., van der Marel, H., Murray, C.A., Le Poole, R.S., Schrijver, H.: 1992, ‘In-orbit performance of the Hipparcos astrometry satellite’, in A&A, 258, pp. 1-6 Perryman, M.A.C., van Leeuwen, F. (eds): 1995, Future possibilities for astrometry in space, ESA, SP-379 Perryman, M.A.C., Lindegren, L., Kovalevsky, J., Turon, C., Høg, E., Grenon, M., Schrijver, H., et al.: 1995, ‘Parallaxes and the Hertzsprung-Russell diagram from the preliminary Hipparcos solution H30’, in A&A, 304, pp. 69-81 Perryman, M.A.C., Lindegren, L., Kovalevsky, J., Høg, E., Bastian, U., Bernacca, P.L., Crézé, M., Donati, F., Grenon, M., Grewing, M., van Leeuwen, F., van der Marel, H., Mignard, F., Murray, C.A., Le Poole, R.S., Schrijver, H., Turon, C., Arenou, F., Frœschlé, M., Petersen, C.S.: 1997a, ‘The Hipparcos Catalogue’, in A&A, 323, pp. L49-L52 Perryman, M.A.C, Brown, A.G.A., Lebreton, Y., Gómez, A., Turon, C., Cayrel de Strobel, G., Mermilliod, J.C., Robichon, N., Kovalevsky, J., Crifo, F.: 1997b, ‘The Hyades: distance, structure, dynamics, and age’, in A&A, in press Perryman, M.A.C., Lindegren, L., Turon, C.: 1997c, ‘The scientific goals of the GAIA mission’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 743-748 Perryman, M.A.C., Bernacca, P.-L. (eds): 1997, ‘Hipparcos, Venice’97’, ESA, SP-402 Platais, I., Girard, T.M., van Altena, W.F., Ma, W.Z., Lindegren, L., Crifo, F., Jahreiß, H.: 1995, ‘A study of systematic positional errors in the SPM plates’, in A&A, 304, pp. 141-149 Provencal, J.L., Shipman, H.L., Høg, E., Thejll, P.: 1997, ‘Testing the white dwarf mass-radius relation with Hipparcos’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 375-378 Quist, C.F., Lindegren, L., Söderhjelm, S.: 1997, ‘Using Hipparcos transit data for aperture synthesis imaging’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 257-262 Ratier, G., Bertaux, J.L., Langen, J., Paulsen, T., Popescu, A., Spiero, F.: 1997, ‘Hipparcos as input catalogue for the GOMOS ENVISAT instrument’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 195-198 Renson, P.: 1980, ‘Méthode de recherche des périodes des étoiles variables’, in A&A, 63, pp. 125-129 THE HIPPARCOS MISSION 407 Réquième, Y., Morrison, L.V., Helmer, L., Fabricius, C., Lindegren, L., Frœschlé, M., van Leeuwen, F., Mignard, F., Perryman, M.A.C, Turon, C.: 1995, ‘Meridian circle reductions using preliminary Hipparcos positions’, in A&A, 304, pp. 121-126 Réquième, Y., Le Campion, J.F., Montignac, G., Daigne, G., Mazurier, J.M., Rapaport, M., Viateau, B., Benevides-Soares, P., Teixeira, R., Høg, E., Makarov, V.V.: 1997, ‘CCD meridian circle reductions using Tycho positions as reference’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 135-138 Reynolds, J.E., Jauncey, D.L., Staveley-Smith, L., Tzioumis, A.K., de Vegt, C., Zacharias, N., Perryman, M.A.C., van Leeuwen, F., King, E.A., McCulloch, P.M., Russell, J.L., Johnston, K.J., Hindsley, R., Malin, D.F., Argue, A.N., Manchester, R.N., Kesteven, M.J., White, G.L., Jones, P.A.: 1995, ‘Accurate registration of radio and optical images of SN 1987A’ in A&A, 304, pp. 116120 Robichon, N., Turon, C., Makarov, V.V., Perryman, M.A.C., Høg, E., Frœschlé, M., Evans, D.W., Guibert, J., Le Poole, R.S., van Leeuwen, F., Bastian, U., Halbwachs, J.L.: 1995, ‘Schmidt plate astrometric reductions using preliminary Hipparcos and Tycho data’, in A&A, 304, pp. 132-140 Röser, S., Bastian, U., de Boer, K.S., Høg, E., Röser, H.P., Schalinski, C., Schilbach, E., de Vegt, C., Wagner, S.: 1997, ‘DIVA – Towards microarcsecond global astrometry’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 777-782 Rybka, S.P., Yatsenko, A.I.: 1997, ‘GPM1 – a catalogue of absolute proper motions of stars with respect to galaxies’, in A&AS, 121, pp. 243-246 Sahade, J.: 1980, ‘The system of Lyrae’, in Space Sci.Rev., 26, pp. 349-389 Scales, D.R., Snijders, M.A.J., Andreasen, G.K., Grenon, M., Høg, E., van Leeuwen, F., Lindegren, L., Mauder, H.: 1992, ‘Tycho photometry calibration and first results’, in A&A, 258, pp. 211-216 Scargle, J.D.: 1982, ‘Studies in astronomical time series analysis. II. Statistical aspects of spectral analysis of unevenly spaced data’, in ApJ, 263, pp. 835-853 Scargle, J.D.: 1989, ‘Studies in astronomical time series analysis. III. Fourier transform, auto correlation functions and cross-corelation functions of unevenly spaced data’, in ApJ, 343, pp. 874-887 Schmidt-Kaler, Th.: 1982, ‘Physical parameters of stars’, in Landolt-Börnstein, Gruppe VI, Band 2b, p453 Schrijver, J.: 1985, ‘First-look processing’, in The second FAST thinkshop, J.Kovalevsky (ed.), pp. 375378 Schwarzenberg-Czerny, A.: 1989, ‘On the advantage of using analysis of variance in period search’, in MNRAS, 241, pp. 153-165 Söderhjelm, S.: 1987, ‘NDAC simulations of double and multiple stars: detectability and astrometric accuracy’, in The third FAST thinkshop, P.L. Bernacca, J. Kovalevsky (eds.), Padova, pp. 155-162 Söderhjelm, S., Evans, D.W., van Leeuwen, F., Lindegren, L.: 1992, ‘Detection and measurement of double stars with the Hipparcos satellite: NDAC reductions’, in A&A, 258, pp. 157-164 Söderhjelm, S., Lindegren, L., Perryman, M.A.C.: 1997, ‘Binary star masses from Hipparcos’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 251-256 Spence, C.B.: 1978, ‘Environmental torques’, in Spacecraft attitude determination and control, J.R. Wertz (ed.), Reidel, Dordrecht, pp. 566-576 Stellingwerf, R.F.: 1978, ‘Period determination using phase dispersion minimization’, in ApJ, 224, pp. 953-960 Stone, R.C.: 1997, ‘Systematic errors in the FK5 catalogue as determined from CCD observations in the extragalactic reference frame’, in AJ, in press Straižys, V., Høg, E.: 1995, ‘An optimum eight-colour photometric system for a survey satellite’, in Future possibilities for astrometry in space, M.A.C. Perryman, F. van Leeuwen (eds.), ESA, SP-379, pp. 191-196 Straižys, V., Høg, E., Davis Philip, A.G.: 1997, ‘The Strömvil seven-colour photometric system for GAIA: classification of peculiar stars’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 761-765 Terlevich, E.: 1987, ‘Evolution of N -body open clusters’, in MNRAS, 224, pp. 193-225 Tian, K.P., van Leeuwen, F., Zhao, J.L., Su, C.G.: 1996, ‘Proper motions of stars in the region of the Orion Nebula cluster (C 0532-054)’, in A&AS, 118, pp. 503-515 408 F. VAN LEEUWEN Tommasini Montanari, T., Bucciarelli, B., Tramontin, C.: 1985, ‘Two different methods for the solution of the large least squares problem in the sphere reconstruction’, in The second FAST thinkshop, J. Kovalevsky (ed.), Marseille, pp. 243-254 Torres, G., Stefanik, R.P., Latham, D.W.: 1997a, ‘The Hyades binary 51 Tauri: Spectroscopic detection of the primary, the distance to the cluster, and the mass-luminosity relation’, in ApJ, 474, pp. 257271 Torres, G., Stefanik, R.P., Latham, D.W.: 1997b, ‘The Hyades binary Finsen 342 (70 Tauri): A doublelined spectroscopic orbit, the distance to the cluster, and the mass-luminosity relation’, in ApJ, 479, pp. 268-278 Torres, G., Stefanik, R.P., Latham, D.W.: 1997c, ‘The Hyades binaries 1 Tauri and 2 Tauri: The distance to the cluster, and the mass-luminosity relation’, in ApJ, 485, pp. 167-181 Tucholke, H.-J., Brosche, P., Odenkirchen, M.: 1997, ‘The Bonn contribution to the extragalactic link of the Hipparcos proper motion system’, in A&AS, 122, pp. 433-440 Turon, C., Gómez, A., Crifo, F., Crézé, M., Perryman, M.A.C., Morin, D., Arenou, F., Nicolet, B., Charleton, M., Egret, D.: 1992a, ‘The Hipparcos Input Catalogue. I. Star selection’, in A&A, 258, pp. 74-81 Turon, C., Arenou, F., Evans, D.W., van Leeuwen, F.: 1992b, ‘Comparisons of the first results from the Hipparcos star mappers with the Hipparcos Input Catalogue’, in A&A, 258, pp. 125-133 Turon, C., van Leeuwen, F.: 1995, ‘Hipparcos data on pulsating stars’, in Astrophysical applications of stellar pulsation, ASP conference series, 83, R.S. Stobie, P.A. Whitelock (eds.), pp. 241-251 Turon, C., Réquième, Y., Grenon, M., Gómez, A., Morin, D., Crifo, F., Arenou, F., Frœschlé, M., Mignard, F., Perryman, M.A.C., Argue, A.N., Crézé, M., Dommanget, J., Egret, D., Jahreiß, H., Mennessier, M.-O., Pham, A., Evans, D.W., van Leeuwen, F., Lindegren, L.: 1995, ‘Properties of the Hipparcos Input Catalogue’, in A&A, 304, pp. 82-93 Udry, S., Mayor, M., Andersen, J., Crifo, F., Grenon, M., Imbert, M., Lindegren, L., Maurice, E., Nordström, B., Pernier, B., Prevot, L., Traversa, G., Turon, C.: 1997, ‘Coravel radial velocity surveys of late-type stars of the Hipparcos mission’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 693-698 Upgren, A.R.: 1977, ‘Internal and external errors in trigonometric parallaxes’, in Vistas in Astr., 21, pp. 241-264 Vaccari, E., Delaney, W.: 1983, ‘Statistical test for recognizing double star signals’, in The FAST thinkshop: Processing of the scientific data from the ESA astrometric satellite Hipparcos, P.L. Bernacca (ed.), Padova, pp. 221-223 Vasilevskis, S., van Leeuwen, F., Nicholson, W., Murray, C.A.: 1979, ‘Internal motions in the central field of the Pleiades’, in A&AS, 37, pp. 333-343 Vauclair, G., Schmidt, H., Koester, D., Allard, N.: 1997, ‘White dwarfs observed by Hipparcos’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 371-374 Viateau, B., Rapaport, M.: 1997, ‘Improvement of the orbits of astroids and the mass of (1) Ceres’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 91-94 Vondrák, J.: 1996, ‘Indirect linking of the Hipparcos catalog to extragalactic reference frame via Earth orientation parameters’, in Dynamics, ephemerides and astrometry in the solar system, S. Ferraz-Mello, B. Morando, J.E. Arlot (eds.), IAU Symp. 172, Dordrecht: Kluwer Academic Publishers, pp. 491-496 Vondrák, J., Ron, C., Pešek, I., Čepek, A.: 1997, ‘The Hipparcos catalogue: a reference frame for Earth orientation in 1899.7–1992.0’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 95-100 Waldrop, M.M.: 1989, ‘Hipparcos: In the low-orbit blues’, in Science, Vol.245, p. 808 Whitelock, P.A., van Leeuwen, F., Feast, M.W.: 1997, ‘The luminosities and diameters of Mira variables from Hipparcos parallaxes’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 213-217 Wicenec, A., Bässgen, G.: 1992, ‘Tycho background determination and monitoring’, in A&A, 258, pp. 206-210 Wicenec, A., van Leeuwen, F.: 1995, ‘The Tycho star mapper background analysis’, in A&A, 304, pp. 160-167 THE HIPPARCOS MISSION 409 Wielen, R., Jahreiß, H.,, Krüger, R.: 1983, ‘The determination of the luminosity function of nearby stars’, in The nearby stars and the stellar luminosity function, IAU coll.76, A.G. Davis Philip, A.R. Upgren (eds.), L.Davis Press, Schenectady, NY, pp. 163-169 Wielen, R.: 1996, ‘Principles of statistical astrometry’, in Astr.Rechen Inst.Heidelberg, preprint series, Preprint No. 65 Wielen, R., Schwan, H., Dettbarn, C., Jahreiß, H., Lenhardt, H.: 1997, ‘Statistical Astrometry based on a comparison of individual proper motions and positions of stars in the FK5 and in the Hipparcos catalogue’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 727-732 Wilcox, J.Z., Wilcox, T.J.: 1995, ‘Algorithm for the extraction of periodic signals from sparse, irregularly sampled data’, in it A&AS, 112, pp. 153-165 Yoshizawa, M., Sato, K., Nishikawa, J., Fukushima, T., Miyamoto, M.: 1997, ‘An optical/infrared astrometric satellite project LIGHT’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 795-797 Zacharias, N., de Vegt, C., Winter, L., Johnston, K.J.: 1995, ‘A radio-optical reference frame. VIII. CCD observations from KPNO and CTIO: Internal calibration and first results’, in AJ, 110, pp. 3093-3106 Zacharias, N.: 1997, ‘Astrometric quality of the USNO CCD astrograph (UCA)’, in AJ, 113, pp. 19251932 Zacharias, N., de Vegt, C., Murray, C.A.: 1997, ‘CPC2 plate reductions with Hipparcos stars: first results’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 8590 de Zeeuw, P.T., Brown, A.G.A, de Bruijne, J.H.J., Hoogerwerf, R., Lub, J., Le Poole, R.S., Blaauw, A.: 1997, ‘Structure and evolution of nearby OB associations’, in Hipparcos, Venice’97, M.A.C. Perryman, P.-L. Bernacca (eds.), ESA, SP-402, pp. 495-500