Binary mass function

Last updated

In astronomy, the binary mass function or simply mass function is a function that constrains the mass of the unseen component (typically a star or exoplanet) in a single-lined spectroscopic binary star or in a planetary system. It can be calculated from observable quantities only, namely the orbital period of the binary system, and the peak radial velocity of the observed star. The velocity of one binary component and the orbital period provide information on the separation and gravitational force between the two components, and hence on the masses of the components.

Contents

Introduction

Two bodies orbiting a common center of mass, indicated by the red plus. The larger body has a higher mass, and therefore a smaller orbit and a lower orbital velocity than its lower-mass companion. Orbit2.gif
Two bodies orbiting a common center of mass, indicated by the red plus. The larger body has a higher mass, and therefore a smaller orbit and a lower orbital velocity than its lower-mass companion.

The binary mass function follows from Kepler's third law when the radial velocity of one binary component is known. [1] Kepler's third law describes the motion of two bodies orbiting a common center of mass. It relates the orbital period with the orbital separation between the two bodies, and the sum of their masses. For a given orbital separation, a higher total system mass implies higher orbital velocities. On the other hand, for a given system mass, a longer orbital period implies a larger separation and lower orbital velocities.

Because the orbital period and orbital velocities in the binary system are related to the masses of the binary components, measuring these parameters provides some information about the masses of one or both components. [2] However, the true orbital velocity is often unknown, because velocities in the plane of the sky are much more difficult to determine than velocities along the line of sight. [1]

Radial velocity is the velocity component of orbital velocity in the line of sight of the observer. Unlike true orbital velocity, radial velocity can be determined from Doppler spectroscopy of spectral lines in the light of a star, [3] or from variations in the arrival times of pulses from a radio pulsar. [4] A binary system is called a single-lined spectroscopic binary if the radial motion of only one of the two binary components can be measured. In this case, a lower limit on the mass of the other, unseen component can be determined. [1]

The true mass and true orbital velocity cannot be determined from the radial velocity because the orbital inclination is generally unknown. (The inclination is the orientation of the orbit from the point of view of the observer, and relates true and radial velocity. [1] ) This causes a degeneracy between mass and inclination. [5] [6] For example, if the measured radial velocity is low, this can mean that the true orbital velocity is low (implying low mass objects) and the inclination high (the orbit is seen edge-on), or that the true velocity is high (implying high mass objects) but the inclination low (the orbit is seen face-on).

Derivation for a circular orbit

Radial velocity curve with peak radial velocity K=1 m/s and orbital period 2 years. Doppler Shift vs Time.svg
Radial velocity curve with peak radial velocity K=1 m/s and orbital period 2 years.

The peak radial velocity is the semi-amplitude of the radial velocity curve, as shown in the figure. The orbital period is found from the periodicity in the radial velocity curve. These are the two observable quantities needed to calculate the binary mass function. [2]

The observed object of which the radial velocity can be measured is taken to be object 1 in this article, its unseen companion is object 2.

Let and be the stellar masses, with the total mass of the binary system, and the orbital velocities, and and the distances of the objects to the center of mass. is the semi-major axis (orbital separation) of the binary system.

We start out with Kepler's third law, with the orbital frequency and the gravitational constant,

Using the definition of the center of mass location, , [1] we can write

Inserting this expression for into Kepler's third law, we find

which can be rewritten to

The peak radial velocity of object 1, , depends on the orbital inclination (an inclination of 0° corresponds to an orbit seen face-on, an inclination of 90° corresponds to an orbit seen edge-on). For a circular orbit (orbital eccentricity = 0) it is given by [7]

After substituting we obtain

The binary mass function (with unit of mass) is [8] [7] [2] [9] [1] [6] [10]

For an estimated or assumed mass of the observed object 1, a minimum mass can be determined for the unseen object 2 by assuming . The true mass depends on the orbital inclination. The inclination is typically not known, but to some extent it can be determined from observed eclipses, [2] be constrained from the non-observation of eclipses, [8] [9] or be modelled using ellipsoidal variations (the non-spherical shape of a star in binary system leads to variations in brightness over the course of an orbit that depend on the system's inclination). [11]

Limits

In the case of (for example, when the unseen object is an exoplanet [8] ), the mass function simplifies to

In the other extreme, when (for example, when the unseen object is a high-mass black hole), the mass function becomes [2]

and since for , the mass function gives a lower limit on the mass of the unseen object 2. [6]

In general, for any or ,

Eccentric orbit

In an orbit with eccentricity , the mass function is given by [7] [12]

Applications

X-ray binaries

If the accretor in an X-ray binary has a minimum mass that significantly exceeds the Tolman–Oppenheimer–Volkoff limit (the maximum possible mass for a neutron star), it is expected to be a black hole. This is the case in Cygnus X-1, for example, where the radial velocity of the companion star has been measured. [13] [14]

Exoplanets

An exoplanet causes its host star to move in a small orbit around the center of mass of the star-planet system. This 'wobble' can be observed if the radial velocity of the star is sufficiently high. This is the radial velocity method of detecting exoplanets. [5] [3] Using the mass function and the radial velocity of the host star, the minimum mass of an exoplanet can be determined. [15] [16] :9 [12] [17] Applying this method on Proxima Centauri, the closest star to the solar system, led to the discovery of Proxima Centauri b, a terrestrial planet with a minimum mass of 1.27 M🜨. [18]

Pulsar planets

Pulsar planets are planets orbiting pulsars, and several have been discovered using pulsar timing. The radial velocity variations of the pulsar follow from the varying intervals between the arrival times of the pulses. [4] The first exoplanets were discovered this way in 1992 around the millisecond pulsar PSR 1257+12. [19] Another example is PSR J1719-1438, a millisecond pulsar whose companion, PSR J1719-1438 b, has a minimum mass approximate equal to the mass of Jupiter, according to the mass function. [8]

Related Research Articles

<span class="mw-page-title-main">Acceleration</span> Rate of change of velocity

In mechanics, acceleration is the rate of change of the velocity of an object with respect to time. Acceleration is one of several components of kinematics, the study of motion. Accelerations are vector quantities. The orientation of an object's acceleration is given by the orientation of the net force acting on that object. The magnitude of an object's acceleration, as described by Newton's Second Law, is the combined effect of two causes:

<span class="mw-page-title-main">Centripetal force</span> Force directed to the center of rotation

A centripetal force is a force that makes a body follow a curved path. The direction of the centripetal force is always orthogonal to the motion of the body and towards the fixed point of the instantaneous center of curvature of the path. Isaac Newton described it as "a force by which bodies are drawn or impelled, or in any way tend, towards a point as to a centre". In the theory of Newtonian mechanics, gravity provides the centripetal force causing astronomical orbits.

In classical mechanics, a harmonic oscillator is a system that, when displaced from its equilibrium position, experiences a restoring force F proportional to the displacement x:

<span class="mw-page-title-main">Spherical coordinate system</span> 3-dimensional coordinate system

In mathematics, a spherical coordinate system is a coordinate system for three-dimensional space where the position of a given point in space is specified by three numbers: the radial distancer connecting the point to the fixed point of origin—located on a fixed polar axis, or z-axis; and the polar angle θ of the radial line r; and the azimuthal angle φ of the radial line r. The polar angle θ is measured between the z-axis and the radial line r. The azimuthal angle φ is measured between the orthogonal projection of the radial line r onto the reference x-y-plane—which is orthogonal to the z-axis and passes through the fixed point of origin)— and either of the fixed x-axis or y-axis, both of which are orthogonal to the z-axis and to each other.

A visual binary is a gravitationally bound binary star system that can be resolved into two stars. These stars are estimated, via Kepler's third law, to have periods ranging from a few years to thousands of years. A visual binary consists of two stars, usually of a different brightness. Because of this, the brighter star is called the primary and the fainter one is called the companion. If the primary is too bright, relative to the companion, this can cause a glare making it difficult to resolve the two components. However, it is possible to resolve the system if observations of the brighter star show it to wobble about a centre of mass. In general, a visual binary can be resolved into two stars with a telescope if their centres are separated by a value greater than or equal to one arcsecond, but with modern professional telescopes, interferometry, or space-based equipment, stars can be resolved at closer distances.

In Newtonian physics, free fall is any motion of a body where gravity is the only force acting upon it. In the context of general relativity, where gravitation is reduced to a space-time curvature, a body in free fall has no force acting on it.

<span class="mw-page-title-main">Radial velocity</span> Velocity of an object as the rate of distance change between the object and a point

The radial velocity or line-of-sight velocity of a target with respect to an observer is the rate of change of the vector displacement between the two points. It is formulated as the vector projection of the target-observer relative velocity onto the relative direction or line-of-sight (LOS) connecting the two points.

In physics, circular motion is a movement of an object along the circumference of a circle or rotation along a circular arc. It can be uniform, with a constant rate of rotation and constant tangential speed, or non-uniform with a changing rate of rotation. The rotation around a fixed axis of a three-dimensional body involves the circular motion of its parts. The equations of motion describe the movement of the center of mass of a body, which remains at a constant distance from the axis of rotation. In circular motion, the distance between the body and a fixed point on its surface remains the same, i.e., the body is assumed rigid.

<span class="mw-page-title-main">Orbital decay</span> Process that leads to gradual decrease of the distance between two orbiting bodies

Orbital decay is a gradual decrease of the distance between two orbiting bodies at their closest approach over many orbital periods. These orbiting bodies can be a planet and its satellite, a star and any object orbiting it, or components of any binary system. If left unchecked, the decay eventually results in termination of the orbit when the smaller object strikes the surface of the primary; or for objects where the primary has an atmosphere, the smaller object burns, explodes, or otherwise breaks up in the larger object's atmosphere; or for objects where the primary is a star, ends with incineration by the star's radiation. Collisions of stellar-mass objects are usually accompanied by effects such as gamma-ray bursts and detectable gravitational waves.

<span class="mw-page-title-main">Projectile motion</span> Motion of launched objects due to gravity


Projectile motion is a form of motion experienced by an object or particle that is projected in a gravitational field, such as from Earth's surface, and moves along a curved path under the action of gravity only. In the particular case of projectile motion on Earth, most calculations assume the effects of air resistance are passive and negligible. The curved path of objects in projectile motion was shown by Galileo to be a parabola, but may also be a straight line in the special case when it is thrown directly upward or downward. The study of such motions is called ballistics, and such a trajectory is a ballistic trajectory. The only force of mathematical significance that is actively exerted on the object is gravity, which acts downward, thus imparting to the object a downward acceleration towards the Earth’s center of mass. Because of the object's inertia, no external force is needed to maintain the horizontal velocity component of the object's motion. Taking other forces into account, such as aerodynamic drag or internal propulsion, requires additional analysis. A ballistic missile is a missile only guided during the relatively brief initial powered phase of flight, and whose remaining course is governed by the laws of classical mechanics.

<span class="mw-page-title-main">Spacecraft flight dynamics</span> Application of mechanical dynamics to model the flight of space vehicles

Spacecraft flight dynamics is the application of mechanical dynamics to model how the external forces acting on a space vehicle or spacecraft determine its flight path. These forces are primarily of three types: propulsive force provided by the vehicle's engines; gravitational force exerted by the Earth and other celestial bodies; and aerodynamic lift and drag.

<span class="mw-page-title-main">Methods of detecting exoplanets</span>

Any planet is an extremely faint light source compared to its parent star. For example, a star like the Sun is about a billion times as bright as the reflected light from any of the planets orbiting it. In addition to the intrinsic difficulty of detecting such a faint light source, the light from the parent star causes a glare that washes it out. For those reasons, very few of the exoplanets reported as of April 2014 have been observed directly, with even fewer being resolved from their host star.

<span class="mw-page-title-main">Doppler spectroscopy</span> Indirect method for finding extrasolar planets and brown dwarfs

Doppler spectroscopy is an indirect method for finding extrasolar planets and brown dwarfs from radial-velocity measurements via observation of Doppler shifts in the spectrum of the planet's parent star. As of November 2022, about 19.5% of known extrasolar planets have been discovered using Doppler spectroscopy.

The two-body problem in general relativity is the determination of the motion and gravitational field of two bodies as described by the field equations of general relativity. Solving the Kepler problem is essential to calculate the bending of light by gravity and the motion of a planet orbiting its sun. Solutions are also used to describe the motion of binary stars around each other, and estimate their gradual loss of energy through gravitational radiation.

<span class="mw-page-title-main">Stellar rotation</span> Angular motion of a star about its axis

Stellar rotation is the angular motion of a star about its axis. The rate of rotation can be measured from the spectrum of the star, or by timing the movements of active features on the surface.

HD 114762 b is a small red dwarf star, in the HD 114762 system, formerly thought to be a massive gaseous extrasolar planet, approximately 126 light-years (38.6 pc) away in the constellation of Coma Berenices. This optically undetected companion to the late F-type main-sequence star HD 114762 was discovered in 1989 by Latham, et al., and confirmed in an October 1991 paper by Cochran, et al. It was thought to be the first discovered exoplanet

<span class="mw-page-title-main">Minimum mass</span> Lowest possible mass of the celestial object

In astronomy, minimum mass is the lower-bound calculated mass of observed objects such as planets, stars and binary systems, nebulae, and black holes.

<span class="mw-page-title-main">Kepler orbit</span> Celestial orbit whose trajectory is a conic section in the orbital plane

In celestial mechanics, a Kepler orbit is the motion of one body relative to another, as an ellipse, parabola, or hyperbola, which forms a two-dimensional orbital plane in three-dimensional space. A Kepler orbit can also form a straight line. It considers only the point-like gravitational attraction of two bodies, neglecting perturbations due to gravitational interactions with other objects, atmospheric drag, solar radiation pressure, a non-spherical central body, and so on. It is thus said to be a solution of a special case of the two-body problem, known as the Kepler problem. As a theory in classical mechanics, it also does not take into account the effects of general relativity. Keplerian orbits can be parametrized into six orbital elements in various ways.

In classical mechanics, the central-force problem is to determine the motion of a particle in a single central potential field. A central force is a force that points from the particle directly towards a fixed point in space, the center, and whose magnitude only depends on the distance of the object to the center. In a few important cases, the problem can be solved analytically, i.e., in terms of well-studied functions such as trigonometric functions.

In astrophysics the chirp mass of a compact binary system determines the leading-order orbital evolution of the system as a result of energy loss from emitting gravitational waves. Because the gravitational wave frequency is determined by orbital frequency, the chirp mass also determines the frequency evolution of the gravitational wave signal emitted during a binary's inspiral phase. In gravitational wave data analysis it is easier to measure the chirp mass than the two component masses alone.

References

  1. 1 2 3 4 5 6 Karttunen, Hannu; Kröger, Pekka; Oja, Heikki; Poutanen, Markku & Donner, Karl J., eds. (2007) [1st pub. 1987]. "Chapter 9: Binary Stars and Stellar Masses". Fundamental Astronomy. Springer Verlag. pp. 221–227. ISBN   978-3-540-34143-7.
  2. 1 2 3 4 5 Podsiadlowski, Philipp. "The Evolution of Binary Systems, in Accretion Processes in Astrophysics" (PDF). Cambridge University Press . Retrieved April 20, 2016.
  3. 1 2 "Radial Velocity – The First Method that Worked". The Planetary Society . Retrieved April 20, 2016.
  4. 1 2 "The Binary Pulsar PSR 1913+16". Cornell University . Retrieved April 26, 2016.
  5. 1 2 Brown, Robert A. (2015). "True Masses of Radial-Velocity Exoplanets". The Astrophysical Journal . 805 (2): 188. arXiv: 1501.02673 . Bibcode:2015ApJ...805..188B. doi:10.1088/0004-637X/805/2/188. S2CID   119294767.
  6. 1 2 3 Larson, Shane. "Binary Stars" (PDF). Utah State University. Archived from the original (PDF) on April 12, 2015. Retrieved April 26, 2016.
  7. 1 2 3 Tauris, T.M. & van den Heuvel, E.P.J. (2006). "Chapter 16: Formation and evolution of compact stellar X-ray sources". In Lewin, Walter & van der Klis, Michiel (eds.). Compact stellar X-ray sources . Cambridge, UK: Cambridge University Press. pp.  623–665. arXiv: astro-ph/0303456 . ISBN   978-0-521-82659-4.
  8. 1 2 3 4 Bailes, M.; Bates, S. D.; Bhalerao, V.; Bhat, N. D. R.; Burgay, M.; Burke-Spolaor, S.; d'Amico, N.; Johnston, S.; et al. (2011). "Transformation of a Star into a Planet in a Millisecond Pulsar Binary". Science . 333 (6050): 1717–1720. arXiv: 1108.5201 . Bibcode:2011Sci...333.1717B. doi:10.1126/science.1208890. PMID   21868629. S2CID   206535504.
  9. 1 2 van Kerkwijk, M. H.; Breton, R. P.; Kulkarni, S. R. (2011). "Evidence for a Massive Neutron Star from a Radial-velocity Study of the Companion to the Black-widow Pulsar PSR B1957+20". The Astrophysical Journal . 728 (2): 95. arXiv: 1009.5427 . Bibcode:2011ApJ...728...95V. doi:10.1088/0004-637X/728/2/95. S2CID   37759376.
  10. "Binary Mass Function". COSMOS – The SAO Encyclopedia of Astronomy, Swinburne University of Technology . Retrieved April 20, 2016.
  11. "The Orbital Inclination". Yale University. Archived from the original on May 14, 2020. Retrieved February 17, 2017.
  12. 1 2 Boffin, H. M. J. (2012). "The mass-ratio distribution of spectroscopic binaries". In Arenou, F. & Hestroffer, D. (eds.). Proceedings of the workshop "Orbital Couples: Pas de Deux in the Solar System and the Milky Way". pp. 41–44. Bibcode:2012ocpd.conf...41B. ISBN   978-2-910015-64-0.{{cite book}}: |journal= ignored (help)
  13. Mauder, H. (1973), "On the Mass Limit of the X-ray Source in Cygnus X-1", Astronomy and Astrophysics , 28: 473–475, Bibcode:1973A&A....28..473M
  14. "Observational Evidence for Black Holes" (PDF). University of Tennessee. Archived from the original (PDF) on October 10, 2017. Retrieved November 3, 2016.
  15. "Documentation and Methodology". Exoplanet Data Explorer . Retrieved April 25, 2016.
  16. Butler, R.P.; Wright, J. T.; Marcy, G. W.; Fischer, D. A.; Vogt, S. S.; Tinney, C. G.; Jones, H. R. A.; Carter, B. D.; et al. (2006). "Catalog of Nearby Exoplanets". The Astrophysical Journal . 646 (1): 505–522. arXiv: astro-ph/0607493 . Bibcode:2006ApJ...646..505B. doi:10.1086/504701. S2CID   119067572.
  17. Kolena, John. "Detecting Invisible Objects: a guide to the discovery of Extrasolar Planets and Black Holes". Duke University . Retrieved April 25, 2016.
  18. Anglada-Escudé, Guillem; Amado, Pedro J.; Barnes, John; et al. (2016). "A terrestrial planet candidate in a temperate orbit around Proxima Centauri". Nature. 536 (7617): 437–440. arXiv: 1609.03449 . Bibcode:2016Natur.536..437A. doi:10.1038/nature19106. PMID   27558064. S2CID   4451513.
  19. Wolszczan, D. A.; Frail, D. (9 January 1992). "A planetary system around the millisecond pulsar PSR1257+12". Nature . 355 (6356): 145–147. Bibcode:1992Natur.355..145W. doi:10.1038/355145a0. S2CID   4260368.