Visual binary

Last updated

A visual binary is a gravitationally bound binary star system [1] 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. [2] However, it is possible to resolve the system if observations of the brighter star show it to wobble about a centre of mass. [3] 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.

Contents

For a visual binary system, measurements taken need to specify, in arc-seconds, the apparent angular separation on the sky and the position angle  which is the angle measured eastward from North in degrees  of the companion star relative to the primary star. Taken over a period of time, the apparent relative orbit of the visual binary system will appear on the celestial sphere. The study of visual binaries reveals useful stellar characteristics: masses, densities, surface temperatures, luminosity, and rotation rates. [4]

Distance

In order to work out the masses of the components of a visual binary system, the distance to the system must first be determined, since from this astronomers can estimate the period of revolution and the separation between the two stars. The trigonometric parallax provides a direct method of calculating a star's mass. This will not apply to the visual binary systems, but it does form the basis of an indirect method called the dynamical parallax. [5]

Trigonometric parallax

In order to use this method of calculating distance, two measurements are made of a star, one each at opposite sides of the Earth's orbit about the Sun. The star's position relative to the more distant background stars will appear displaced. The parallax value is considered to be the displacement in each direction from the mean position, equivalent to the angular displacement from observations one astronomical unit apart. The distance , in parsecs is found from the following equation,

Where is the parallax, measured in units of arc-seconds. [6]

Dynamical parallax

This method is used solely for binary systems. The mass of the binary system is assumed to be twice that of the Sun. Kepler's Laws are then applied and the separation between the stars is determined. Once this distance is found, the distance away can be found via the arc subtended in the sky, providing a temporary distance measurement. From this measurement and the apparent magnitudes of both stars, the luminosities can be found, and by using the mass–luminosity relationship, the masses of each star. These masses are used to re-calculate the separation distance, and the process is repeated a number of times, with accuracies as high as 5% being achieved. A more sophisticated calculation factors in a star's loss of mass over time. [5]

Spectroscopic parallax

Spectroscopic parallax is another commonly used method for determining the distance to a binary system. No parallax is measured, the word is simply used to place emphasis on the fact that the distance is being estimated. In this method, the luminosity of a star is estimated from its spectrum. It is important to note that the spectra from distant stars of a given type are assumed to be the same as the spectra of nearby stars of the same type. The star is then assigned a position on the Hertzsprung-Russel diagram based on where it is in its life-cycle. The star's luminosity can be estimated by comparison of the spectrum of a nearby star. The distance is then determined via the following inverse square law:

where is the apparent brightness and is the luminosity.

Using the Sun as a reference we can write

where the subscript represents a parameter associated with the Sun.

Rearranging for gives an estimate for the distance. [7]

Kepler's laws

The two stars orbiting each other, as well as their centre of mass, must obey Kepler's laws. This means that the orbit is an ellipse with the centre of mass at one of the two foci (Kepler's 1st law) and the orbital motion satisfies the fact that a line joining the star to the centre of mass sweeps out equal areas over equal time intervals (Kepler's 2nd law). The orbital motion must also satisfy Kepler's 3rd law. [8]

Kepler's 3rd Law can be stated as follows: "The square of the orbital period of a planet is directly proportional to the cube of its semi-major axis." Mathematically, this translates as

where is the orbital period of the planet and is the semi-major axis of the orbit. [8]

Newton's generalisation

Consider a binary star system. This consists of two objects, of mass and , orbiting around their centre of mass. has position vector and orbital velocity , and has position vector and orbital velocity relative to the centre of mass. The separation between the two stars is denoted , and is assumed to be constant. Since the gravitational force acts along a line joining the centers of both stars, we can assume the stars have an equivalent time period around their center of mass, and therefore a constant separation between each other. [9]

To arrive at Newton's version of Kepler's 3rd law we can start by considering Newton's 2nd law which states: "The net force acting on an object is proportional to the objects mass and resultant acceleration."

where is the net force acting on the object of mass , and is the acceleration of the object. [10]

Applying the definition of centripetal acceleration to Newton's second law gives a force of

[11]

Then using the fact that the orbital velocity is given as

[11]

we can state the force on each star as

and

If we apply Newton's 3rd law- "For every action there is an equal and opposite reaction"

[10]

We can set the force on each star equal to each other.

This reduces to

If we assume that the masses are not equal, then this equation tells us that the smaller mass remains farther from the centre of mass than does the larger mass.

The separation of the two objects is

Since and would form a line starting from opposite directions and joining at the centre of mass.

Now we can substitute this expression into one of the equations describing the force on the stars and rearrange for to find an expression relating the position of one star to the masses of both and the separation between them. Equally, this could have been solved for . We find that

Substituting this equation into the equation for the force on one of the stars, setting it equal to Newton's Universal Law of Gravitation (namely, , [10] and solving for the period squared yields the required result.

[10]

This is Newton's version of Kepler's 3rd Law. Unless is in non-standard units, this will not work if mass is measured in solar masses, the orbital period is measured in years, and the orbital semi-major axis is measured in astronomical units (e.g., use the Earth's orbital parameters). It will work if SI units, for instance, are used throughout.

Determining stellar masses

Binary systems are particularly important here  because they are orbiting each other, their gravitational interaction can be studied by observing parameters of their orbit around each other and the centre of mass. Before applying Kepler's 3rd Law, the inclination of the orbit of the visual binary must be taken into account. Relative to an observer on Earth, the orbital plane will usually be tilted. If it is at 0° the planes will be seen to coincide and if at 90° they will be seen edge on. Due to this inclination, the elliptical true orbit will project an elliptical apparent orbit onto the plane of the sky. Kepler's 3rd law still holds but with a constant of proportionality that changes with respect to the elliptical apparent orbit. [12] The inclination of the orbit can be determined by measuring the separation between the primary star and the apparent focus. Once this information is known the true eccentricity and the true semi-major axis can be calculated since the apparent orbit will be shorter than the true orbit, assuming an inclination greater than 0°, and this effect can be corrected for using simple geometry

Where is the true semi-major axis and is the parallax.

Once the true orbit is known, Kepler's 3rd law can be applied. We re-write it in terms of the observable quantities such that

From this equation we obtain the sum of the masses involved in the binary system. Remembering a previous equation we derived,

where

we can solve the ratio of the semi-major axis and therefore a ratio for the two masses since

and

The individual masses of the stars follow from these ratios and knowing the separation between each star and the centre of mass of the system. [4]

Mass–luminosity relationship

In order to find the luminosity of the stars, the rate of flow of radiant energy, otherwise known as radiant flux, must be observed. When the observed luminosities and masses are graphed, the mass–luminosity relation is obtained. This relationship was found by Arthur Eddington in 1924.

Where L is the luminosity of the star and M is its mass. L and M are the luminosity and mass of the Sun. [13] The value  = 3.5 is commonly used for main-sequence stars. [14] This equation and the usual value of a = 3.5 only applies to main-sequence stars with masses 2M < M < 20M and does not apply to red giants or white dwarfs. For these stars, the equation applies with different constants, since these stars have different masses. For the different ranges of masses, an adequate form of the Mass–Luminosity Relation is

The greater a star's luminosity, the greater its mass will be. The absolute magnitude or luminosity of a star can be found by knowing the distance to it and its apparent magnitude. The stars bolometric magnitude is plotted against its mass, in units of the Sun's mass. This is determined through observation and then the mass of the star is read of the plot. Giants and main sequence stars tend to agree with this, but super giants do not and neither do white dwarfs. The Mass–Luminosity Relation is very useful because, due to the observation of binaries, particularly the visual binaries since the masses of many stars have been found this way, astronomers have gained insight into the evolution of stars, including how they are born. [5] [13] [15]

Spectral classification

Generally speaking, there are three classes of binary systems. These can be determined by considering the colours of the two components.

"1. Systems consisting of a red or reddish primary star and a blueish secondary star, usually a magnitude or more fainter... 2. Systems in which the differences in magnitude and colour are both small... 3. Systems in which the fainter star is the redder of the two..."

The luminosity of class 1. binaries is greater than that of class 3. binaries. There is a relationship between the colour difference of binaries and their reduced proper motions. In 1921, Frederick C. Leonard, at the Lick Observatory, wrote "1. The spectrum of the secondary component of a dwarf star is generally redder than that of the primary, whereas the spectrum of the fainter component of a giant star is usually bluer than that of the brighter one. In both cases, the absolute difference in spectral class seems ordinarily to be related to the disparity between the components...2. With some exceptions, the spectra of the components of double stars are so related to each other that they conform to the Hertzsprung-Russell configuration of the stars..."

An interesting case for visual binaries occurs when one or both components are located above or below the Main-Sequence. If a star is more luminous than a Main-Sequence star, it is either very young, and therefore contracting due to gravity, or is at the post Main-Sequence stage of its evolution. The study of binaries is useful here because, unlike with single stars, it is possible to determine which reason is the case. If the primary is gravitationally contracting, then the companion will be further away from the Main-Sequence than the primary since the more massive star becomes a Main-Sequence star much faster than the less massive star. [16]

Related Research Articles

<span class="mw-page-title-main">Luminosity</span> Measurement of radiant electromagnetic power emitted by an object

Luminosity is an absolute measure of radiated electromagnetic power (light), the radiant power emitted by a light-emitting object over time. In astronomy, luminosity is the total amount of electromagnetic energy emitted per unit of time by a star, galaxy, or other astronomical objects.

The Eddington luminosity, also referred to as the Eddington limit, is the maximum luminosity a body can achieve when there is balance between the force of radiation acting outward and the gravitational force acting inward. The state of balance is called hydrostatic equilibrium. When a star exceeds the Eddington luminosity, it will initiate a very intense radiation-driven stellar wind from its outer layers. Since most massive stars have luminosities far below the Eddington luminosity, their winds are mostly driven by the less intense line absorption. The Eddington limit is invoked to explain the observed luminosity of accreting black holes such as quasars.

<span class="mw-page-title-main">Two-body problem</span> Motion problem in classical mechanics

In classical mechanics, the two-body problem is to predict the motion of two massive objects which are abstractly viewed as point particles. The problem assumes that the two objects interact only with one another; the only force affecting each object arises from the other one, and all other objects are ignored.

<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">Elliptic orbit</span> Kepler orbit with an eccentricity of less than one

In astrodynamics or celestial mechanics, an elliptic orbit or elliptical orbit is a Kepler orbit with an eccentricity of less than 1; this includes the special case of a circular orbit, with eccentricity equal to 0. In a stricter sense, it is a Kepler orbit with the eccentricity greater than 0 and less than 1. In a wider sense, it is a Kepler orbit with negative energy. This includes the radial elliptic orbit, with eccentricity equal to 1.

<span class="mw-page-title-main">Radiation zone</span>

A radiation zone, or radiative region is a layer of a star's interior where energy is primarily transported toward the exterior by means of radiative diffusion and thermal conduction, rather than by convection. Energy travels through the radiation zone in the form of electromagnetic radiation as photons.

In orbital mechanics, mean motion is the angular speed required for a body to complete one orbit, assuming constant speed in a circular orbit which completes in the same time as the variable speed, elliptical orbit of the actual body. The concept applies equally well to a small body revolving about a large, massive primary body or to two relatively same-sized bodies revolving about a common center of mass. While nominally a mean, and theoretically so in the case of two-body motion, in practice the mean motion is not typically an average over time for the orbits of real bodies, which only approximate the two-body assumption. It is rather the instantaneous value which satisfies the above conditions as calculated from the current gravitational and geometric circumstances of the body's constantly-changing, perturbed orbit.

In classical mechanics, the Kepler problem is a special case of the two-body problem, in which the two bodies interact by a central force F that varies in strength as the inverse square of the distance r between them. The force may be either attractive or repulsive. The problem is to find the position or speed of the two bodies over time given their masses, positions, and velocities. Using classical mechanics, the solution can be expressed as a Kepler orbit using six orbital elements.

<span class="mw-page-title-main">Gliese 667</span> Triple star system in the constellation Scorpius

Gliese 667 is a triple-star system in the constellation Scorpius lying at a distance of about 7.2 parsecs from Earth. All three of the stars have masses smaller than the Sun. There is a 12th-magnitude star close to the other three, but it is not gravitationally bound to the system. To the naked eye, the system appears to be a single faint star of magnitude 5.89.

<span class="mw-page-title-main">Initial mass function</span>

In astronomy, the initial mass function (IMF) is an empirical function that describes the initial distribution of masses for a population of stars during star formation. IMF not only describes the formation and evolution of individual stars, it also serves as an important link that describes the formation and evolution of galaxies. The IMF is often given as a probability density function (PDF) that describes the probability of a star that has a certain mass. It differs from the present-day mass function (PDMF), which describes the current distribution of masses of stars, such as red giants, white dwarfs, neutron stars, and black holes, after a period of time of evolution away from the main sequence stars. IMF is derived from the luminosity function while PDMF is derived from the present-day luminosity function. IMF and PDMF can be linked through the "stellar creation function". Stellar creation function is defined as the number of stars per unit volume of space in a mass range and a time interval. For all the main sequence stars have greater lifetimes than the galaxy, IMF and PDMF are equivalent. Similarly, IMF and PDMF are equivalent in brown dwarfs due to their unlimited lifetimes.

<span class="mw-page-title-main">Hayashi track</span> Luminosity–temperature relationship in stars

The Hayashi track is a luminosity–temperature relationship obeyed by infant stars of less than 3 M in the pre-main-sequence phase of stellar evolution. It is named after Japanese astrophysicist Chushiro Hayashi. On the Hertzsprung–Russell diagram, which plots luminosity against temperature, the track is a nearly vertical curve. After a protostar ends its phase of rapid contraction and becomes a T Tauri star, it is extremely luminous. The star continues to contract, but much more slowly. While slowly contracting, the star follows the Hayashi track downwards, becoming several times less luminous but staying at roughly the same surface temperature, until either a radiative zone develops, at which point the star starts following the Henyey track, or nuclear fusion begins, marking its entry onto the main sequence.

In astronomy, the distance to a visual binary star may be estimated from the masses of its two components, the size of their orbit, and the period of their orbit about one another. A dynamical parallax is an (annual) parallax which is computed from such an estimated distance.

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">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.

<span class="mw-page-title-main">Semi-major and semi-minor axes</span> Term in geometry; longest and shortest semidiameters of an ellipse

In geometry, the major axis of an ellipse is its longest diameter: a line segment that runs through the center and both foci, with ends at the two most widely separated points of the perimeter. The semi-major axis is the longest semidiameter or one half of the major axis, and thus runs from the centre, through a focus, and to the perimeter. The semi-minor axis of an ellipse or hyperbola is a line segment that is at right angles with the semi-major axis and has one end at the center of the conic section. For the special case of a circle, the lengths of the semi-axes are both equal to the radius of the circle.

In astrophysics, the mass–luminosity relation is an equation giving the relationship between a star's mass and its luminosity, first noted by Jakob Karl Ernst Halm. The relationship is represented by the equation:

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.

HD 98649 is a G-type yellow dwarf star, classified as a G4V, that has approximately the same mass and diameter as the Sun, but has only 86% of its luminosity. It is considered a solar analog. HD 98649 is about 138 light-years from earth. HD 98649 is found in the Crater constellation.

In astronomy, the binary mass function or simply mass function is a function that constrains the mass of the unseen component 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.

<span class="mw-page-title-main">Kepler-1229b</span> Super-Earth orbiting Kepler-1229

Kepler-1229b is a confirmed super-Earth exoplanet, likely rocky, orbiting within the habitable zone of the red dwarf Kepler-1229, located about 870 light years from Earth in the constellation of Cygnus. It was discovered in 2016 by the Kepler space telescope. The exoplanet was found by using the transit method, in which the dimming effect that a planet causes as it crosses in front of its star is measured.

References

  1. Argyle, R. W. (2012), Observing and Measuring Visual Double Stars, The Patrick Moore Practical Astronomy Series, Springer Science & Business Media, pp. 71–75, ISBN   978-1461439455
  2. The Binary Stars, Robert Grant Aitken, New York: Dover, 1964, p. 41.
  3. "Binary Systems and Stellar Parameters" (PDF). Archived from the original (PDF) on 2013-11-04. Retrieved 2013-11-02.
  4. 1 2 Michael Zeilik; Stephan A. Gregory & Elske V. P. Smith (1998). Introductory Astronomy and Astrophysics. Brooks/Cole. ISBN   978-0030062285.
  5. 1 2 3 Mullaney, James (2005). Double and multiple stars and how to observe them . Springer. p.  27. ISBN   1-85233-751-6. Mass–Luminosity relation distance binary.
  6. Martin Harwit (20 April 2000). Astrophysical Concepts. Springer. ISBN   0-387-94943-7.
  7. European Space Agency, Stellar distances
  8. 1 2 Leonard Susskind & George Hrabovsky (2013). The Theoretical Minimum: What You Need To Know To Start Doing Physics . the Penguin Group. ISBN   978-1846147982.
  9. "The Physics of Binary Stars" . Retrieved 2013-10-15.
  10. 1 2 3 4 Bradley W. Carroll & Dale A. Ostlie (2013). An Introduction to Modern Astrophysics. Pearson. ISBN   978-1292022932.
  11. 1 2 Hugh D. Young (2010). University Physics. Bertrams. ISBN   978-0321501301.
  12. "Kepler's laws, Binaries, and Stellar Masses" (PDF). Retrieved 2013-11-04.
  13. 1 2 Salaris, Maurizio; Santi Cassisi (2005). Evolution of stars and stellar populations. John Wiley & Sons. pp. 138–140. ISBN   0-470-09220-3.
  14. "Mass–luminosity relationship". Hyperphysics. Retrieved 2009-08-23.
  15. Duric, Nebojsa (2004). Advanced astrophysics. Cambridge University Press. p. 19. ISBN   978-0-521-52571-8.
  16. William P. Bidelman, "Spectral Classifications of Visual Binaries having Primaries above the Main Sequence", Lick Observatory, University of California, Retrieved 24/11/13