Light year and cosmic scales. How long to fly to the nearest star? (8 photos) Distance to a star 20 light

And how many potentially explosive stars are located at an unsafe distance?

A supernova is an incredible explosion of a star - and almost beyond the limits of human imagination. If our Sun were to explode as a supernova, the resulting shock wave would probably not destroy the entire Earth, but the side of the Earth facing the Sun would disappear. Scientists believe that the temperature of the planet as a whole would increase by about 15 times. Moreover, the Earth will not remain in orbit.

A sudden decrease in the mass of the Sun could free the planet and send it wandering into space. It is clear that the distance to the Sun - 8 light minutes - is not safe. Fortunately, our Sun is not a star destined to explode in a supernova. But other stars, outside our solar system, might. What is the closest safe distance? Scientific literature shows 50 to 100 light years as the closest safe distance between Earth and a supernova.


Optical image of the remnant of supernova 1987A, taken by the Hubble Space Telescope

What happens if a supernova explodes near the Earth? Let's consider the explosion of a star other than our Sun, but still at an unsafe distance. Let's say a supernova is 30 light years away. Dr. Mark Reid, senior astronomer at the Harvard-Smithsonian Center for Astrophysics, says:

“... if there was a supernova that was about 30 light years away, it would lead to strong impacts on the Earth, possibly mass extinctions. X-rays and more energetic gamma rays from a supernova can destroy the ozone layer that protects us from the sun's ultraviolet rays. It could also ionize nitrogen and oxygen in the atmosphere, leading to the formation of large amounts of nitrous oxide-like smog in the atmosphere."

Moreover, if a supernova were to explode 30 light years away, phytoplankton and reef communities would be particularly affected. Such an event severely depletes the base of the ocean food chain.

Let's assume that the explosion was a little more distant. The explosion of a nearby star could leave Earth, its surface, and ocean life relatively untouched. But any relatively close explosion would still "douse" us with gamma rays and other high-energy particles. This radiation can cause mutations in earthly life. In addition, the radiation of the nearest supernova could change our climate.

It is known that a supernova has not exploded at such a close distance in the known history of mankind. The most recent supernova visible to the eye was supernova 1987A, in 1987. It was about 168,000 light years away. Prior to this, the last flash visible to the eye was recorded by Johannes Kepler in 1604. Approximately 20,000 light years away, it shone brighter than any star in the night sky. This explosion was visible even in daylight! As far as we know, this did not cause noticeable consequences.

How many potential supernovae are closer to us than 50 to 100 light years away? The answer depends on the type of supernova. A Type II supernova is an aging massive star that is collapsing. There are no stars massive enough to do this within 50 light-years of Earth.

But there are also Type I supernovae - caused by the collapse of a small, pale white dwarf star. These stars are dim and hard to spot, so we can't be sure how many there are around. Probably several hundred of these stars are within 50 light years.

Relative sizes of IK Pegasi A (left), B (bottom, center) and the Sun (right).

The star IK Pegasi B is the closest candidate for the role of a supernova prototype. It is part of a binary star system located about 150 light years from our Sun and solar system.

The main star in the system, IK Pegasi A, is an ordinary main sequence star not unlike our Sun. A potential Type I supernova is another star, IK Pegasi B, a massive white dwarf that is extremely small and dense. When star A begins to evolve into a red giant, it is expected to grow to a radius where it will collide with a white dwarf or start pulling matter from A's expanded gas envelope. When star B becomes massive enough, it can explode as a supernova.

What about Betelgeuse? Another star often mentioned in the history of supernovae is Betelgeuse, one of the brightest stars in our sky, part of the famous constellation of Orion. Betelgeuse is a supergiant star. It is inherently very bright.

However, this brilliance comes at a price. Betelgeuse is one of the most famous stars in the sky because it will explode someday. The enormous energy of Betelgeuse requires that the fuel be used up quickly (relatively speaking), and in fact Betelgeuse is already nearing the end of its life. Someday soon (astronomically speaking) it will run out of fuel and then experience a spectacular Type II supernova explosion. When this happens, Betelgeuse will become brighter for weeks or months, perhaps as bright as the full moon and visible in broad daylight.

When will it happen? Probably not in our lives, but no one knows for sure. It could be tomorrow or a million years in the future. When this happens, everyone on Earth will witness an impressive event in the night sky, but earthly life will not be affected. This is because Betelgeuse is 430 light years away.

How often do supernovae break out in our galaxy? No one knows. Scientists have suggested that high-energy supernova radiation has already caused mutations in terrestrial species, maybe even humans.

According to one estimate, there may be one dangerous supernova event in the vicinity of the Earth every 15 million years. Other scientists say that, on average, a supernova explosion occurs within 10 parsecs (33 light years) of Earth every 240 million years. So you see what we really don't know. But you can compare these numbers to a few million years, the time when humans are considered to have existed on the planet, and four and a half billion years for the very age of the Earth.

And, if you do, you will see that a supernova is sure to explode near the Earth - but probably not in the foreseeable future of mankind.

Like( 3 ) I do not like( 0 )

Stars are the most common type of celestial bodies in the universe. There are about 6000 stars up to the 6th magnitude, about a million up to the 11th magnitude, and about 2 billion of them in the entire sky up to the 21st magnitude.

All of them, like the Sun, are hot self-luminous gas balls, in the depths of which huge energy is released. However, the stars, even in the most powerful telescopes, are visible as luminous points, since they are very far from us.

1. Annual parallax and distances to stars

The radius of the Earth turns out to be too small to serve as a basis for measuring the parallactic displacement of stars and for determining the distances to them. Even in the time of Copernicus, it was clear that if the Earth really revolves around the Sun, then the apparent positions of the stars in the sky must change. In six months, the Earth moves by the diameter of its orbit. Directions to the star from opposite points of this orbit must be different. In other words, the stars should have a noticeable annual parallax (Fig. 72).

The annual parallax of a star ρ is the angle at which one could see the semi-major axis of the earth's orbit (equal to 1 AU) from a star if it is perpendicular to the line of sight.

The greater the distance D to the star, the smaller its parallax. The parallactic shift of the star's position in the sky during the year occurs along a small ellipse or circle if the star is at the ecliptic pole (see Fig. 72).

Copernicus tried but failed to detect the parallax of the stars. He correctly asserted that the stars were too far from the Earth for the then existing instruments to detect their parallactic displacement.

The first reliable measurement of the annual parallax of the star Vega was made in 1837 by the Russian academician V. Ya. Struve. Almost simultaneously with him in other countries, the parallaxes of two more stars were determined, one of which was α Centauri. This star, which is not visible in the USSR, turned out to be the closest to us, its annual parallax is ρ = 0.75". At this angle, a wire 1 mm thick is visible to the naked eye from a distance of 280 m. small angular displacements.

Distance to the star where a is the semi-major axis of the earth's orbit. At small angles if p is expressed in arcseconds. Then, taking a = 1 a. e., we get:


Distance to the nearest star α Centauri D \u003d 206 265 ": 0.75" \u003d 270,000 a. e. Light travels this distance in 4 years, while it takes only 8 minutes from the Sun to the Earth, and about 1 s from the Moon.

The distance light travels in a year is called a light year.. This unit is used to measure distance along with the parsec (pc).

A parsec is the distance from which the semi-major axis of the earth's orbit, perpendicular to the line of sight, is visible at an angle of 1".

The distance in parsecs is equal to the reciprocal of the annual parallax, expressed in arcseconds. For example, the distance to the star α Centauri is 0.75" (3/4"), or 4/3 pc.

1 parsec = 3.26 light years = 206,265 AU e. = 3 * 10 13 km.

At present, the measurement of the annual parallax is the main method for determining the distances to stars. Parallaxes have already been measured for very many stars.

By measuring the annual parallax, one can reliably determine the distance to stars located no further than 100 pc, or 300 light years.

Why is it not possible to accurately measure the annual parallax of more than o distant stars?

The distance to more distant stars is currently determined by other methods (see §25.1).

2. Apparent and absolute magnitude

The luminosity of the stars. After astronomers were able to determine the distances to the stars, it was found that the stars differ in apparent brightness, not only because of the difference in their distance, but also because of the difference in their luminosity.

The luminosity of a star L is the power of emission of light energy in comparison with the power of emission of light by the Sun.

If two stars have the same luminosity, then the star that is farthest from us has a lower apparent brightness. Comparing stars by luminosity is possible only if their apparent brightness (magnitude) is calculated for the same standard distance. Such a distance in astronomy is considered to be 10 pc.

The apparent stellar magnitude that a star would have if it were at a standard distance D 0 \u003d 10 pc from us was called the absolute magnitude M.

Let us consider the quantitative ratio of the apparent and absolute stellar magnitudes of a star at a known distance D to it (or its parallax p). Recall first that a difference of 5 magnitudes corresponds to a brightness difference of exactly 100 times. Consequently, the difference in apparent stellar magnitudes of two sources is equal to one, when one of them is brighter than the other exactly one time (this value is approximately equal to 2.512). The brighter the source, the smaller its apparent magnitude is considered. In the general case, the ratio of the apparent brightness of any two stars I 1:I 2 is related to the difference in their apparent magnitudes m 1 and m 2 by a simple relationship:


Let m be the apparent magnitude of a star located at a distance D. If it were observed from a distance D 0 = 10 pc, its apparent magnitude m 0 would, by definition, be equal to the absolute magnitude M. Then its apparent brightness would change by

At the same time, it is known that the apparent brightness of a star varies inversely with the square of its distance. So

(2)

Hence,

(3)

Taking the logarithm of this expression, we find:

(4)

where p is expressed in arcseconds.

These formulas give the absolute magnitude M from the known apparent magnitude m at a real distance to the star D. From a distance of 10 pc, our Sun would look approximately like a star of the 5th apparent magnitude, i.e. for the Sun M ≈5.

Knowing the absolute magnitude M of a star, it is easy to calculate its luminosity L. Taking the luminosity of the Sun L = 1, by definition of luminosity, we can write that

The values ​​of M and L in different units express the radiation power of the star.

The study of stars shows that they can differ in luminosity by tens of billions of times. In stellar magnitudes, this difference reaches 26 units.

Absolute values stars of very high luminosity are negative and reach M = -9. Such stars are called giants and supergiants. The radiation of the star S Doradus is 500,000 times more powerful than the radiation of our Sun, its luminosity is L=500,000, dwarfs with M=+17 (L=0.000013) have the lowest radiation power.

To understand the reasons for the significant differences in the luminosity of stars, it is necessary to consider their other characteristics, which can be determined on the basis of radiation analysis.

3. Color, spectra and temperature of stars

During your observations, you noticed that the stars have a different color, which is clearly visible in the brightest of them. The color of a heated body, including stars, depends on its temperature. This makes it possible to determine the temperature of stars from the distribution of energy in their continuous spectrum.

The color and spectrum of stars are related to their temperature. In relatively cold stars, radiation in the red region of the spectrum predominates, which is why they have a reddish color. The temperature of red stars is low. It rises sequentially as it goes from red to orange, then to yellow, yellowish, white, and bluish. The spectra of stars are extremely diverse. They are divided into classes, denoted by Latin letters and numbers (see back flyleaf). In the spectra of cool red stars of class M with a temperature of about 3000 K, absorption bands of the simplest diatomic molecules, most often titanium oxide, are visible. The spectra of other red stars are dominated by oxides of carbon or zirconium. Red stars of the first magnitude class M - Antares, Betelgeuse.

In the spectra of yellow G stars, which include the Sun (with a temperature of 6000 K on the surface), thin lines of metals predominate: iron, calcium, sodium, etc. A star like the Sun in terms of spectrum, color and temperature is the bright Chapel in the constellation Auriga.

In the spectra of white class A stars, like Sirius, Vega and Deneb, the hydrogen lines are the strongest. There are many weak lines of ionized metals. The temperature of such stars is about 10,000 K.

In the spectra of the hottest, bluish stars with a temperature of about 30,000 K, lines of neutral and ionized helium are visible.

The temperatures of most stars are between 3,000 and 30,000 K. A few stars have temperatures around 100,000 K.

Thus, the spectra of stars are very different from each other, and they can be used to determine the chemical composition and temperature of the atmospheres of stars. The study of the spectra showed that hydrogen and helium are predominant in the atmospheres of all stars.

The differences in stellar spectra are explained not so much by the diversity of their chemical composition as by the difference in temperature and other physical conditions in stellar atmospheres. At high temperatures, molecules break down into atoms. At an even higher temperature, less durable atoms are destroyed, they turn into ions, losing electrons. Ionized atoms of many chemical elements, like neutral atoms, emit and absorb energy of certain wavelengths. By comparing the intensity of the absorption lines of atoms and ions of the same chemical element, their relative number is theoretically determined. It is a function of temperature. So, from the dark lines of the spectra of stars, you can determine the temperature of their atmospheres.

Stars of the same temperature and color, but different luminosities, have the same spectra in general, but one can notice differences in the relative intensities of some lines. This is due to the fact that at the same temperature the pressure in their atmospheres is different. For example, in the atmospheres of giant stars, the pressure is less, they are rarer. If this dependence is expressed graphically, then the absolute magnitude of the star can be found from the intensity of the lines, and then, using formula (4), the distance to it can be determined.

Problem solution example

Task. What is the luminosity of the star ζ Scorpio, if its apparent magnitude is 3, and the distance to it is 7500 sv. years?


Exercise 20

1. How many times is Sirius brighter than Aldebaran? Is the sun brighter than Sirius?

2. One star is 16 times brighter than the other. What is the difference between their magnitudes?

3. The parallax of Vega is 0.11". How long does it take the light from it to reach the Earth?

4. How many years would it take to fly towards the constellation Lyra at a speed of 30 km / s for Vega to become twice as close?

5. How many times is a star of magnitude 3.4 fainter than Sirius, which has an apparent magnitude of -1.6? What are the absolute magnitudes of these stars if the distance to both is 3 pc?

6. Name the color of each of the stars in Appendix IV according to their spectral type.

How to determine the distance to the stars? How do you know that Alpha Centauri is about 4 light years away? Indeed, by the brightness of a star, as such, you can hardly determine anything - the brilliance of a dim close and bright distant stars can be the same. And yet there are many fairly reliable ways to determine the distance from the Earth to the farthest corners of the universe. Astrometric satellite "Hipparchus" for 4 years of work determined the distances to 118 thousand SPL stars

Whatever physicists say about the three-dimensionality, six-dimensionality or even eleven-dimensionality of space, for the astronomer the observable Universe is always two-dimensional. What is happening in the Cosmos is seen by us as a projection onto the celestial sphere, just as in a movie the whole complexity of life is projected onto a flat screen. On the screen, we can easily distinguish the far from the near thanks to familiarity with the three-dimensional original, but in the two-dimensional scattering of stars there is no visual clue that allows us to turn it into a three-dimensional map suitable for plotting the course of an interstellar ship. Meanwhile, distances are the key to almost half of all astrophysics. How can one distinguish a nearby dim star from a distant but bright quasar without them? Only knowing the distance to an object, one can evaluate its energy, and from here a direct road to understanding its physical nature.

A recent example of the uncertainty of cosmic distances is the problem of sources of gamma-ray bursts, short pulses of hard radiation that come to Earth from various directions about once a day. Initial estimates of their remoteness ranged from hundreds of astronomical units (tens of light hours) to hundreds of millions of light years. Accordingly, the scatter in the models was also impressive - from the annihilation of comets from antimatter on the outskirts of the solar system to the explosions of neutron stars shaking the entire Universe and the birth of white holes. By the mid-1990s, more than a hundred different explanations for the nature of gamma-ray bursts had been proposed. Now, when we were able to estimate the distances to their sources, there are only two models left.

But how to measure the distance if neither the ruler nor the locator beam can reach the object? The triangulation method, widely used in conventional terrestrial geodesy, comes to the rescue. We select a segment of known length - the base, measure from its ends the angles under which a point is visible, which is inaccessible for one reason or another, and then simple trigonometric formulas give the desired distance. When we move from one end of the base to the other, the apparent direction to the point changes, it shifts against the background of distant objects. This is called parallax shift, or parallax. Its value is the smaller, the farther the object is, and the larger, the longer the base.

To measure the distances to the stars, one has to take the maximum base available to astronomers, equal to the diameter of the earth's orbit. The corresponding parallactic displacement of stars in the sky (strictly speaking, half of it) came to be called the annual parallax. Tycho Brahe tried to measure it, who did not like the Copernican idea about the rotation of the Earth around the Sun, and he decided to check it - after all, parallaxes also prove the orbital movement of the Earth. The measurements carried out had an accuracy that was impressive for the 16th century - about one minute of arc, but this was completely insufficient for measuring parallaxes, which Brahe himself had no idea about and concluded that the Copernican system was incorrect.

The distance to star clusters is determined by the main sequence fitting method

The next attack on parallax was made in 1726 by the Englishman James Bradley, the future director of the Greenwich Observatory. At first, it seemed that luck smiled at him: the star Gamma Draco, chosen for observations, indeed fluctuated around its average position with a span of 20 seconds of arc during the year. However, the direction of this shift was different from that expected for parallaxes, and Bradley soon found the correct explanation: the speed of the Earth's orbit adds up to the speed of light coming from the star, and changes its apparent direction. Similarly, raindrops leave sloping paths on the windows of a bus. This phenomenon, called annual aberration, was the first direct evidence of the Earth moving around the Sun, but had nothing to do with parallaxes.

Only a century later, the accuracy of goniometric instruments reached the required level. In the late 30s of the XIX century, in the words of John Herschel, "the wall that prevented penetration into the stellar Universe was broken almost simultaneously in three places." In 1837, Vasily Yakovlevich Struve (at that time the director of the Derpt Observatory, and later of the Pulkovo Observatory) published the parallax of Vega measured by him - 0.12 arc seconds. The following year, Friedrich Wilhelm Bessel reported that the parallax of the star of the 61st Cygnus is 0.3 ". And a year later, the Scottish astronomer Thomas Henderson, who worked in the Southern Hemisphere at the Cape of Good Hope, measured the parallax in the Alpha Centauri system - 1.16" . True, later it turned out that this value was overestimated by 1.5 times and there is not a single star in the entire sky with a parallax of more than 1 second of arc.

For distances measured by the parallactic method, a special unit of length was introduced - parsec (from parallactic second, pc). One parsec contains 206,265 astronomical units, or 3.26 light years. It is from this distance that the radius of the earth's orbit (1 astronomical unit = 149.5 million kilometers) is visible at an angle of 1 second. To determine the distance to a star in parsecs, one must divide one by its parallax in seconds. For example, to the closest star system to us, Alpha Centauri, 1/0.76 = 1.3 parsecs, or 270,000 astronomical units. A thousand parsecs is called a kiloparsec (kpc), a million parsecs is called a megaparsec (Mpc), a billion is called a gigaparsec (Gpc).

Measuring extremely small angles required technical sophistication and great diligence (Bessel, for example, processed more than 400 individual observations of Cygnus 61), but after the first breakthrough, things got easier. By 1890, the parallaxes of already three dozen stars had been measured, and when photography began to be widely used in astronomy, the accurate measurement of parallaxes was completely put on stream. Parallax measurements are the only method for directly determining the distances to individual stars. However, during ground-based observations, atmospheric interference does not allow the parallax method to measure distances above 100 pc. For the universe, this is not a very large value. (“It's not far, a hundred parsecs,” as Gromozeka said.) Where geometric methods fail, photometric methods come to the rescue.

Geometric records

In recent years, the results of measuring distances to very compact sources of radio emission - masers - have been published more and more often. Their radiation falls on the radio range, which makes it possible to observe them on radio interferometers capable of measuring the coordinates of objects with microsecond accuracy, unattainable in the optical range in which stars are observed. Thanks to masers, trigonometric methods can be applied not only to distant objects in our Galaxy, but also to other galaxies. For example, in 2005 Andreas Brunthaler (Germany) and his colleagues determined the distance to the M33 galaxy (730 kpc) by comparing the angular displacement of masers with the rotation speed of this star system. A year later, Ye Xu (China) and colleagues applied the classical parallax method to "local" maser sources to measure the distance (2 kpc) to one of the spiral arms of our Galaxy. Perhaps, in 1999, J. Hernstin (USA) and colleagues managed to advance the farthest. Tracking the movement of masers in the accretion disk around the black hole at the core of the active galaxy NGC 4258, astronomers have determined that this system is 7.2 Mpc away from us. To date, this is an absolute record of geometric methods.

Astronomers standard candles

The farther away from us is the source of radiation, the dimmer it is. If you know the true luminosity of an object, then by comparing it with the visible brightness, you can find the distance. Probably the first to apply this idea to the measurement of distances to stars was Huygens. At night, he observed Sirius, and during the day he compared its brilliance with a tiny hole in the screen that covered the Sun. Having chosen the size of the hole so that both brightnesses coincided, and comparing the angular values ​​of the hole and the solar disk, Huygens concluded that Sirius is 27,664 times farther from us than the Sun. This is 20 times less than the real distance. The error was partly due to the fact that Sirius is actually much brighter than the Sun, and partly due to the difficulty of comparing brightness from memory.

A breakthrough in the field of photometric methods occurred with the advent of photography in astronomy. At the beginning of the 20th century, the Harvard College Observatory carried out large-scale work to determine the brightness of stars from photographic plates. Particular attention was paid to variable stars, whose brightness fluctuates. Studying variable stars of a special class - Cepheids - in the Small Magellanic Cloud, Henrietta Levitt noticed that the brighter they are, the longer the period of fluctuation of their brightness: stars with a period of several tens of days turned out to be about 40 times brighter than stars with a period of about a day.

Since all Levitt Cepheids were in the same star system - the Small Magellanic Cloud - it could be considered that they were removed from us at the same (albeit unknown) distance. This means that the difference in their apparent brightness is associated with real differences in luminosity. It remained to determine the distance to one Cepheid by a geometric method in order to calibrate the entire dependence and to be able, by measuring the period, to determine the true luminosity of any Cepheid, and from it the distance to the star and the star system containing it.

But, unfortunately, there are no Cepheids in the vicinity of the Earth. The nearest of them, the Polar Star, is, as we now know, 130 pc from the Sun, that is, it is beyond the reach of ground-based parallax measurements. This did not allow to throw a bridge directly from parallaxes to Cepheids, and astronomers had to build a structure, which is now figuratively called the distance ladder.

An intermediate stage on it was open star clusters, including from several tens to hundreds of stars, connected by a common time and place of birth. If you plot the temperature and luminosity of all the stars in the cluster, most of the points will fall on one inclined line (more precisely, a strip), which is called the main sequence. The temperature is determined with high accuracy from the spectrum of the star, and the luminosity is determined from the apparent brightness and distance. If the distance is unknown, the fact again comes to the rescue that all the stars in the cluster are almost the same distance from us, so that within the cluster, the apparent brightness can still be used as a measure of luminosity.

Since the stars are the same everywhere, the main sequences of all clusters must match. The differences are due only to the fact that they are at different distances. If we determine the distance to one of the clusters by a geometric method, then we will find out what the “real” main sequence looks like, and then, by comparing data from other clusters with it, we will determine the distances to them. This technique is called "main sequence fitting". For a long time, the Pleiades and Hyades served as a standard for it, the distances to which were determined by the method of group parallaxes.

Fortunately for astrophysics, Cepheids have been found in about two dozen open clusters. Therefore, by measuring the distances to these clusters by fitting the main sequence, one can "reach the ladder" to the Cepheids, which are on its third step.

As an indicator of distances, Cepheids are very convenient: there are relatively many of them - they can be found in any galaxy and even in any globular cluster, and being giant stars, they are bright enough to measure intergalactic distances from them. Thanks to this, they have earned many high-profile epithets, such as "beacons of the universe" or "mileposts of astrophysics." The Cepheid "ruler" stretches up to 20 Mpc - this is about a hundred times the size of our Galaxy. Further, they can no longer be distinguished even with the most powerful modern instruments, and in order to climb the fourth rung of the distance ladder, you need something brighter.

To the ends of the universe

One of the most powerful extragalactic methods for measuring distances is based on a pattern known as the Tully-Fisher relation: the brighter a spiral galaxy, the faster it rotates. When a galaxy is seen edge-on or at a significant inclination, half of its matter is approaching us due to rotation, and half is receding, which leads to the expansion of spectral lines due to the Doppler effect. This expansion determines the speed of rotation, according to it - the luminosity, and then from a comparison with the apparent brightness - the distance to the galaxy. And, of course, to calibrate this method, galaxies are needed, the distances to which have already been measured using Cepheids. The Tully-Fisher method is very long-range and covers galaxies that are hundreds of megaparsecs away from us, but it also has a limit, since it is not possible to obtain enough high-quality spectra for too distant and faint galaxies.

In a somewhat larger range of distances, another "standard candle" operates - type Ia supernovae. Flashes of such supernovae are "same type" thermonuclear explosions of white dwarfs with a mass slightly higher than the critical one (1.4 solar masses). Therefore, there is no reason for them to vary greatly in power. Observations of such supernovae in nearby galaxies, the distances to which can be determined from Cepheids, seem to confirm this constancy, and therefore cosmic thermonuclear explosions are now widely used to determine distances. They are visible even billions of parsecs from us, but you never know the distance to which galaxy you can measure, because it is not known in advance exactly where the next supernova will break out.

So far, only one method allows moving even further - redshifts. Its history, like the history of Cepheids, begins simultaneously with the 20th century. In 1915, the American Westo Slifer, studying the spectra of galaxies, noticed that in most of them the lines are redshifted relative to the "laboratory" position. In 1924, the German Karl Wirtz noticed that this shift is the stronger, the smaller the angular size of the galaxy. However, only Edwin Hubble in 1929 managed to bring these data into a single picture. According to the Doppler effect, the redshift of the lines in the spectrum means that the object is moving away from us. Comparing the spectra of galaxies with the distances to them, determined by the Cepheids, Hubble formulated the law: the speed of the removal of a galaxy is proportional to the distance to it. The coefficient of proportionality in this ratio is called the Hubble constant.

Thus, the expansion of the Universe was discovered, and with it the possibility of determining the distances to galaxies from their spectra, of course, provided that the Hubble constant is tied to some other “rulers”. Hubble himself performed this binding with an error of almost an order of magnitude, which was corrected only in the mid-1940s, when it became clear that Cepheids are divided into several types with different "period - luminosity" ratios. The calibration was carried out again based on the "classical" Cepheids, and only then the value of the Hubble constant became close to modern estimates: 50-100 km/s for every megaparsec of distance to the galaxy.

Now, redshifts are used to determine distances to galaxies that are thousands of megaparsecs away from us. True, these distances are indicated in megaparsecs only in popular articles. The fact is that they depend on the model of evolution of the Universe adopted in the calculations, and besides, in expanding space it is not entirely clear what distance is meant: the one at which the galaxy was at the moment of emission of radiation, or the one at which it is located at the time of its reception on Earth, or the distance traveled by light on the way from the starting point to the end point. Therefore, astronomers prefer to indicate for distant objects only the directly observed redshift value, without converting it to megaparsecs.

Redshifts are currently the only method for estimating "cosmological" distances comparable to the "size of the Universe", and at the same time, this is perhaps the most widespread technique. In July 2007, a catalog of redshifts of 77,418,767 galaxies was published. True, when creating it, a somewhat simplified automatic technique for analyzing spectra was used, and therefore errors could creep into some values.

Team play

Geometric methods for measuring distances are not limited to annual parallax, in which the apparent angular displacements of stars are compared with the movements of the Earth in its orbit. Another approach relies on the motion of the Sun and stars relative to each other. Imagine a star cluster flying past the Sun. According to the laws of perspective, the visible trajectories of its stars, like rails on the horizon, converge to one point - the radiant. Its position indicates the angle at which the cluster flies to the line of sight. Knowing this angle, one can decompose the motion of the cluster stars into two components - along the line of sight and perpendicular to it along the celestial sphere - and determine the proportion between them. The radial velocity of stars in kilometers per second is measured by the Doppler effect and, taking into account the proportion found, the projection of the velocity onto the sky is calculated - also in kilometers per second. It remains to compare these linear velocities of the stars with the angular velocities determined from the results of long-term observations - and the distance will be known! This method works up to several hundred parsecs, but is applicable only to star clusters and is therefore called the group parallax method. This is how the distances to the Hyades and Pleiades were first measured.

Down the stairs leading up

Building our ladder to the outskirts of the universe, we were silent about the foundation on which it rests. Meanwhile, the parallax method gives the distance not in reference meters, but in astronomical units, that is, in the radii of the earth's orbit, the value of which was also not immediately determined. So let's look back and go down the ladder of cosmic distances to Earth.

Probably the first to determine the remoteness of the Sun was Aristarchus of Samos, who proposed the heliocentric system of the world one and a half thousand years before Copernicus. It turned out that the Sun is 20 times farther from us than the Moon. This estimate, as we now know, underestimated by a factor of 20, lasted until the Kepler era. Although he himself did not measure the astronomical unit, he already noted that the Sun should be much further than Aristarchus (and all other astronomers followed him) thought.

The first more or less acceptable estimate of the distance from the Earth to the Sun was obtained by Jean Dominique Cassini and Jean Richet. In 1672, during the opposition of Mars, they measured its position against the background of the stars simultaneously from Paris (Cassini) and Cayenne (Richet). The distance from France to French Guiana served as the base of the parallactic triangle, from which they determined the distance to Mars, and then calculated the astronomical unit from the equations of celestial mechanics, deriving a value of 140 million kilometers.

Over the next two centuries, the transits of Venus across the solar disk became the main tool for determining the scale of the solar system. By observing them simultaneously from different parts of the globe, it is possible to calculate the distance from the Earth to Venus, and hence all other distances in the solar system. In the XVIII-XIX centuries, this phenomenon was observed four times: in 1761, 1769, 1874 and 1882. These observations became one of the first international scientific projects. Large-scale expeditions were equipped (the English expedition of 1769 was led by the famous James Cook), special observation stations were created ... And if at the end of the 18th century Russia only provided French scientists with the opportunity to observe the passage from its territory (from Tobolsk), then in 1874 and 1882 Russian scientists have already taken an active part in the research. Unfortunately, the exceptional complexity of the observations has led to a significant discrepancy in the estimates of the astronomical unit - from about 147 to 153 million kilometers. A more reliable value - 149.5 million kilometers - was obtained only at the turn of the 19th-20th centuries from observations of asteroids. And, finally, it must be taken into account that the results of all these measurements were based on the knowledge of the length of the base, in the role of which, when measuring the astronomical unit, the radius of the Earth acted. So in the end, the foundation of the ladder of cosmic distances was laid by surveyors.

Only in the second half of the 20th century did fundamentally new methods for determining cosmic distances appear at the disposal of scientists - laser and radar. They made it possible to increase the accuracy of measurements in the solar system hundreds of thousands of times. The error of radar for Mars and Venus is several meters, and the distance to the corner reflectors installed on the Moon is measured to within centimeters. The currently accepted value of the astronomical unit is 149,597,870,691 meters.

The difficult fate of "Hipparchus"

Such a radical progress in the measurement of the astronomical unit raised the question of the distances to stars in a new way. The accuracy of determining parallaxes is limited by the Earth's atmosphere. Therefore, back in the 1960s, the idea arose to bring a goniometric instrument into space. It was realized in 1989 with the launch of the European astrometric satellite Hipparchus. This name is a well-established, although formally not quite correct translation of the English name HIPPARCOS, which is an abbreviation for High Precision Parallax Collecting Satellite (“satellite for collecting high-precision parallaxes”) and does not coincide with the English spelling of the name of the famous ancient Greek astronomer - Hipparchus, the author of the first star directory.

The creators of the satellite set themselves a very ambitious task: to measure the parallaxes of more than 100 thousand stars with millisecond accuracy, that is, to “reach out” to stars located hundreds of parsecs from the Earth. It was necessary to clarify the distances to several open star clusters, in particular the Hyades and the Pleiades. But most importantly, it became possible to "jump over the step" by directly measuring the distances to the Cepheids themselves.

The expedition began with trouble. Due to a failure in the upper stage, the Hipparchus did not enter the calculated geostationary orbit and remained on an intermediate highly elongated trajectory. The specialists of the European Space Agency nevertheless managed to cope with the situation, and the orbital astrometric telescope successfully operated for 4 years. The processing of results lasted the same amount of time, and in 1997 a stellar catalog was published with parallaxes and proper motions of 118,218 luminaries, including about two hundred Cepheids.

Unfortunately, in a number of issues the desired clarity has not yet come. The result for the Pleiades turned out to be the most incomprehensible - it was assumed that Hipparchus would clarify the distance, which was previously estimated at 130-135 parsecs, but in practice it turned out that Hipparchus corrected it, getting a value of only 118 parsecs. Acceptance of the new value would require adjustments to both the theory of stellar evolution and the scale of intergalactic distances. This would be a serious problem for astrophysics, and the distance to the Pleiades began to be carefully checked. By 2004, several groups had independently obtained estimates of the distance to the cluster in the range from 132 to 139 pc. Offensive voices began to be heard with suggestions that the consequences of putting the satellite into the wrong orbit still could not be completely eliminated. Thus, in general, all parallaxes measured by him were called into question.

The Hipparchus team was forced to admit that the measurements were generally accurate, but might need to be re-processed. The point is that parallaxes are not measured directly in space astrometry. Instead, Hipparchus measured the angles between numerous pairs of stars over and over again for four years. These angles change both due to the parallactic displacement and due to the proper motions of the stars in space. To "pull out" exactly the values ​​of parallaxes from observations, a rather complicated mathematical processing is required. This is what I had to repeat. The new results were published at the end of September 2007, but it is not yet clear how much of an improvement this has made.

But the problems of Hipparchus do not end there. The Cepheid parallaxes determined by him turned out to be insufficiently accurate for a confident calibration of the "period-luminosity" ratio. Thus, the satellite failed to solve the second task facing it. Therefore, several new projects of space astrometry are currently being considered in the world. The European Gaia project, which is scheduled to launch in 2012, is the closest to implementation. Its principle of operation is the same as that of the Hipparchus - repeated measurements of the angles between pairs of stars. However, thanks to powerful optics, it will be able to observe much dimmer objects, and the use of the interferometry method will increase the accuracy of measuring angles to tens of microseconds of arc. It is assumed that Gaia will be able to measure kiloparsec distances with an error of no more than 20% and will determine the positions of about a billion objects over several years of work. Thus, a three-dimensional map of a significant part of the Galaxy will be constructed.

Aristotle's universe ended at nine distances from the Earth to the Sun. Copernicus believed that the stars were 1,000 times further away than the Sun. Parallaxes pushed even the nearest stars away by light years. At the very beginning of the 20th century, the American astronomer Harlow Shapley, using Cepheids, determined that the diameter of the Galaxy (which he identified with the Universe) was measured in tens of thousands of light years, and thanks to Hubble, the boundaries of the Universe expanded to several gigaparsecs. How final are they?

Of course, each rung of the distance ladder has its own, larger or smaller errors, but in general, the scales of the Universe are well defined, verified by various methods that are independent of each other, and add up to a single consistent picture. So the current boundaries of the universe seem unshakable. However, this does not mean that one day we will not want to measure the distance from it to some neighboring universe!

Proxima Centauri.

Here's a classic backfill question. Ask your friends Which one is closest to us?" and then watch them list nearest stars. Maybe Sirius? Alpha something there? Betelgeuse? The answer is obvious - it is; a massive ball of plasma located about 150 million kilometers from Earth. Let's clarify the question. Which star is closest to the Sun?

nearest star

You have probably heard that - the third brightest star in the sky at a distance of only 4.37 light years from. But Alpha Centauri not a single star, it is a system of three stars. First, a binary star (binary star) with a common center of gravity and an orbital period of 80 years. Alpha Centauri A is only slightly more massive and brighter than the Sun, while Alpha Centauri B is slightly less massive than the Sun. There is also a third component in this system, a dim red dwarf Proxima Centauri (Proxima Centauri).


Proxima Centauri- That's what it is closest star to our sun, located at a distance of only 4.24 light years.

Proxima Centauri.

Multiple star system Alpha Centauri located in the constellation Centaurus, which is only visible in the southern hemisphere. Unfortunately, even if you see this system, you will not be able to see Proxima Centauri. This star is so dim that you need a powerful enough telescope to see it.

Let's find out the scale of how far Proxima Centauri from U.S. Think about. moves at a speed of almost 60,000 km / h, the fastest in. He overcame this path in 2015 for 9 years. Traveling so fast to get to Proxima Centauri, New Horizons will need 78,000 light years.

Proxima Centauri is the nearest star over 32,000 light years, and it will hold this record for another 33,000 years. It will make its closest approach to the Sun in about 26,700 years, when the distance from this star to the Earth will be only 3.11 light years. In 33,000 years, the nearest star will be Ross 248.

What about the northern hemisphere?

For those of us who live in the northern hemisphere, the nearest visible star is Barnard's Star, another red dwarf in the constellation Ophiuchus (Ophiuchus). Unfortunately, like Proxima Centauri, Barnard's Star is too dim to see with the naked eye.


Barnard's Star.

nearest star, which you can see with the naked eye in the northern hemisphere is Sirius (Alpha Canis Major). Sirius is twice the size and mass of the Sun and is the brightest star in the sky. Located 8.6 light-years away in the constellation Canis Major (Canis Major), it is the most famous star chasing Orion in the night sky during the winter.

How did astronomers measure the distance to stars?

They use a method called . Let's do a little experiment. Hold one arm outstretched at length and place your finger so that some distant object is nearby. Now alternately open and close each eye. Notice how your finger seems to jump back and forth when you look with different eyes. This is the parallax method.

Parallax.

To measure the distance to the stars, you can measure the angle to the star with respect to when the Earth is on one side of the orbit, say in the summer, then 6 months later when the Earth moves to the opposite side of the orbit, and then measure the angle to the star compared to which some distant object. If the star is close to us, this angle can be measured and the distance calculated.

You can really measure the distance in this way to nearby stars, but this method only works up to 100,000 light years.

20 nearest stars

Here is a list of the 20 nearest star systems and their distances in light years. Some of them have several stars, but they are part of the same system.

StarDistance, St. years
Alpha Centauri4,2
Barnard's Star5,9
Wolf 359 (Wolf 359; CN Lion)7,8
Lalande 21185 (Lalande 21185)8,3
Sirius8,6
Leuthen 726-8 (Luyten 726-8)8,7
Ross 154 (Ross 154)9,7
Ross 248 (Ross 24810,3
Epsilon Eridani10,5
Lacaille 9352 (Lacaille 9352)10,7
Ross 128 (Ross 128)10,9
EZ Aquarii (EZ Aquarii)11,3
Procyon (Procyon)11,4
61 Cygni11,4
Struve 2398 (Struve 2398)11,5
Groombridge 34 (Groombridge 34)11,6
Epsilon Indi11,8
DX Cancri11,8
Tau Ceti11,9
GJ 10611,9

According to NASA, there are 45 stars within a radius of 17 light years from the Sun. There are over 200 billion stars in the universe. Some of them are so dim that they are almost impossible to detect. Perhaps with new technologies, scientists will find stars even closer to us.

The title of the article you read "Closest Star to the Sun".

Looking out the train window

The calculation of the distance to the stars did not worry the ancient people much, because in their opinion they were attached to the celestial sphere and were at the same distance from the Earth, which a person could never measure. Where are we, and where are these divine domes?

It took many, many centuries for people to understand: the Universe is somewhat more complicated. To understand the world in which we live, it was necessary to build a spatial model in which each star is removed from us at a certain distance, just as a tourist needs a map to complete a route, not a panoramic photograph of the area.

Parallax, familiar to us from traveling by train or by car, became the first assistant in this complex undertaking. Have you noticed how fast roadside poles flicker against the backdrop of distant mountains? If you noticed, then you can be congratulated: you, unwittingly, discovered an important feature of the parallactic shift - for close objects it is much larger and more noticeable. And vice versa.

What is parallax?

In practice, parallax began to work for a person in geodesy and (where without it ?!) in military affairs. Indeed, who, if not gunners, needs to measure distances to distant objects with the highest possible accuracy? Moreover, the triangulation method is simple, logical and does not require the use of any complex devices. All that is required is to measure two angles and one distance, the so-called base, with acceptable accuracy, and then, using elementary trigonometry, determine the length of one of the legs of a right triangle.

Triangulation in practice

Imagine that you need to determine the distance (d) from one coast to an inaccessible point on the ship. Below we present the algorithm of the necessary actions for this.

  1. Mark two points (A) and (B) on the shore, the distance between which you know (l).
  2. Measure angles α and β.
  3. Calculate d using the formula:

Parallax displacement of loved onesstars against the backdrop of distant

Obviously, the accuracy directly depends on the size of the base: the longer it is, the larger the parallax displacements and angles will be, respectively. For an earthly observer, the maximum possible base is the diameter of the Earth's orbit around the Sun, that is, measurements must be carried out at intervals of six months, when our planet is at the diametrically opposite point of the orbit. Such a parallax is called annual, and the first astronomer who tried to measure it was the famous Dane Tycho Brahe, who became famous for his exceptional scientific pedantry and rejection of the Copernican system.

It is possible that Braga's adherence to the idea of ​​geocentrism played a cruel joke on him: the measured annual parallaxes did not exceed a minute of arc and could well be attributed to instrumental errors. The astronomer with a clear conscience was convinced of the "correctness" of the Ptolemaic system - the Earth does not move anywhere and is located in the center of a small cozy Universe, in which the Sun and other stars are literally within easy reach, only 15-20 times further than the Moon. However, the works of Tycho Brahe were not in vain, becoming the foundation for the discovery of Kepler's laws, which finally put an end to outdated theories of the structure of the solar system.

Star Cartographers

Space "ruler"

It should be noted that, before seriously taking on distant stars, triangulation worked perfectly in our space house. The main task was to determine the distance to the Sun, the same astronomical unit, without the exact knowledge of which the measurements of stellar parallaxes become meaningless. The first person in the world who set himself such a task was the ancient Greek philosopher Aristarchus of Samos, who proposed a heliocentric system of the world 1,500 years before Copernicus. After doing complex calculations based on rather approximate knowledge of that era, he found that the Sun is 20 times further away than the Moon. For many centuries, this value was taken as the truth, becoming one of the basic axioms of the theories of Aristotle and Ptolemy.

Only Kepler, coming close to building a model of the solar system, subjected this value to a serious reassessment. On this scale, it was not possible to connect real astronomical data and the laws of motion of celestial bodies discovered by him. Intuitively, Kepler believed that the Sun was much farther away from the Earth, but, being a theoretician, he did not find a way to confirm (or refute) his conjecture.

It is curious that a correct estimate of the size of an astronomical unit became possible precisely on the basis of Kepler's laws, which set the "rigid" spatial structure of the solar system. Astronomers had its accurate and detailed map, on which it remained only to determine the scale. This is what the Frenchmen Jean Dominique Cassini and Jean Richet did, who measured the position of Mars against the background of distant stars during opposition (in this position, Mars, Earth and the Sun are located on one straight line, and the distance between the planets is minimal).

The measuring points were Paris and the capital of French Guiana, Cayenne, a good 7 thousand kilometers away. The young Richet went to the South American colony, while the venerable Cassini remained "musketeer" in Paris. Upon the return of the young colleague, the scientists sat down to the calculations, and at the end of 1672 they presented the results of their research - according to their calculations, the astronomical unit was equal to 140 million kilometers. Later, to refine the scale of the solar system, astronomers used the transits of Venus across the solar disk, which occurred four times in the 18th-19th centuries. And, perhaps, these studies can be called the first international scientific projects: in addition to England, Germany and France, Russia became an active participant in them. By the beginning of the 20th century, the scale of the solar system was finally established, and the modern value of the astronomical unit was accepted - 149.5 million kilometers.

  1. Aristarchus suggested that the Moon has the shape of a ball and is illuminated by the Sun. Therefore, if the Moon looks "cut" in half, then the Earth-Moon-Sun angle is right.
  2. Aristarchus then calculated the Sun-Earth-Moon angle by direct observation.
  3. Using the rule "the sum of the angles of a triangle is 180 degrees", Aristarchus calculated the angle of the Earth-Sun-Moon.
  4. Applying the ratio of the sides of a right triangle, Aristarchus calculated that the Earth-Moon distance is 20 times greater than the Earth-Sun. Note! Aristarchus did not calculate the exact distance.

Parsecs, parsecs

Cassini and Richet calculated the position of Mars relative to distant stars

And with these initial data it was already possible to claim the accuracy of measurements. In addition, goniometers have reached the desired level. Russian astronomer Vasily Struve, director of the university observatory in the city of Derpt (now Tartu in Estonia), in 1837 published the results of measuring the annual parallax of Vega. It turned out to be equal to 0.12 arc seconds. The baton was picked up by the German Friedrich Wilhelm Bessel, a student of the great Gauss, who a year later measured the parallax of the star 61 in the constellation Cygnus - 0.30 arc seconds, and the Scot Thomas Henderson, who "caught" the famous Alpha Centauri with a parallax of 1.2. Later, however, it turned out that the latter overdid it a little and in fact the star shifts by only 0.7 arc seconds per year.

The accumulated data showed that the annual parallax of stars does not exceed one arc second. It was adopted by scientists to introduce a new unit of measurement - the parsec ("parallactic second" in abbreviation). From such an insane distance by conventional standards, the radius of the earth's orbit is visible at an angle of 1 second. To better visualize the cosmic scale, let's assume that the astronomical unit (and this is the radius of the Earth's orbit, equal to 150 million kilometers) "shrank" into 2 tetrad cells (1 cm). So: you can “see” them at an angle of 1 second ... from two kilometers!

For cosmic depths, a parsec is not a distance, although even light will need three and a quarter years to overcome it. Within just a dozen parsecs, our stellar neighbors can literally be counted on the fingers. When it comes to galactic scales, it is time to operate with kilo- (thousand units) and megaparsecs (respectively, a million), which in our "tetrad" model can already climb into other countries.

A real boom in ultra-precise astronomical measurements began with the advent of photography. "Big-eyed" telescopes with meter lenses, sensitive photographic plates designed for many hours of exposure, precision clock mechanisms that rotate the telescope synchronously with the Earth's rotation - all this made it possible to confidently record annual parallaxes with an accuracy of 0.05 arc seconds and, thus, determine distances up to 100 parsecs. Earth technology is incapable of more (or rather, less) because the capricious and restless earthly atmosphere interferes.

If measurements are taken in orbit, then the accuracy can be significantly improved. It was for this purpose that in 1989 the Hipparcos astrometric satellite (HIPPARCOS, from the English High Precision Parallax Collecting Satellite), developed by the European Space Agency, was launched into low Earth orbit.

  1. As a result of the work of the Hipparchus orbital telescope, a fundamental astrometric catalog was compiled.
  2. With the help of Gaia, a three-dimensional map of a part of our Galaxy was compiled, indicating the coordinates, direction of movement and color of about a billion stars.

The result of his work is a catalog of 120,000 stellar objects with annual parallaxes determined to within 0.01 arc seconds. And its successor, the Gaia satellite (Global Astrometric Interferometer for Astrophysics), launched on December 19, 2013, draws a spatial map of the nearest galactic neighborhood with a billion (!) Objects. And who knows, maybe it will be very useful for our grandchildren.