🎧 New: AI-Generated Podcasts Turn your study notes into engaging audio conversations. Learn more

Loading...
Loading...
Loading...
Loading...
Loading...
Loading...
Loading...

Full Transcript

Option D 16 Astrophysics ESSENTIAL IDEAS One of the most difficult problems in astronomy is coming to terms with the vast distances between stars and galaxies and devising accurate methods for measuring them. A simple diagram that plots the luminosity versus the surface temperature of stars reveals...

Option D 16 Astrophysics ESSENTIAL IDEAS One of the most difficult problems in astronomy is coming to terms with the vast distances between stars and galaxies and devising accurate methods for measuring them. A simple diagram that plots the luminosity versus the surface temperature of stars reveals unusually detailed patterns that help understand the inner workings of stars. Stars follow well-defined patterns from the moment they are created to their eventual death. The Hot Big Bang model is a theory that describes the origin and expansion of the universe and is supported by extensive experimental evidence. The laws of nuclear physics applied to nuclear fusion processes inside stars determine the production of all the elements up to iron. The modern field of cosmology uses advanced experimental and observational techniques to collect data with an unprecedented degree of precision, and as a result very surprising and detailed conclusions about the structure of the universe have been reached. 16.1 (D1: Core) Stellar quantities – one of the most difficult problems in astronomy is coming to terms with the vast distances between stars and galaxies and devising accurate methods for measuring them Nature of Science A topic without practical investigations Astronomy is an unusual topic within the study of physics because the standard ‘scientific method’ is not so obvious. There are no controlled experiments designed to investigate a theory. Instead astronomers make observations and collect data. One consequence of this is that the growth of knowledge in astronomy is very much dependent on the latest technology available to aid observations. No student studying astronomy can fail to be impressed by the depth of knowledge about the universe that astronomers have gained from apparently so little evidence – just the radiation received from outer space! In the first section of this chapter we will begin by summarising what we can see in the night sky and then outline the essential features of stars and stellar systems, before explaining the scale of the universe and the units astronomers use to measure such large distances. Finally we will establish the important relationship between the power emitted from a star and the intensity received here on Earth. Observing the night sky On a clear night, far away from the light pollution of towns, it may be possible to see hundreds of stars in the night sky with the unaided eye. A total of about 5000 stars are visible from Earth with the human eye, but not all can be seen at the same time, or from the same place. What we can see depends on our location, the time of night and the time of year. This variation happens because of the Earth’s motion – its spin on its axis and its orbit around the Sun. In theory, at any one time, in any one place, we might be able to see about half of the visible stars. Stars seem to stay in exactly the same positions/patterns (relative to other stars) over thousands of years and therefore we can locate the stars precisely on a star map, such as shown in Figure 16.1. Although stars are moving very fast, their motion is not usually noticeable from Earth, even over very long periods of time (in human terms) because they are such enormous distances away from us. 2 16 Astrophysics Figure 16.1 A star map for the southern hemisphere Figure 16.2 The apparent rotation of the stars as the Earth spins If you observe the stars over a period of hours on any one night you will notice that they appear to move across the sky from east to west – in exactly the same way as the Sun appears to move during the day. These apparent motions are actually produced because the Earth spins in the opposite direction. Time-lapse photography can be used to show the paths of stars across the sky during the night. Such photographs can even show the complete circular path of stars which are close to the Earth’s extended axis (Figure 16.2). In the course of one day, the Earth’s rotation causes our view of the stars to revolve through 360° but, of course, during the day we are not able to see the stars because of the light from the Sun. (Radio astronomers do not have this problem.) Our night-time view changes slightly from one night to the next and after six months we are looking in exactly the opposite direction, as shown in Figure 16.3. The Sun, the Moon and the five planets that are visible with the unaided eye are all much, much closer to Earth than the stars. Their movements as seen from Earth can seem more complicated and they cannot be located in fixed positions on a star map. The Sun, the Earth, the Moon and the planets all move in approximately the same plane. This means that they follow similar paths across the sky as seen by us as the Earth rotates. Sun In June we look this way to see the night sky In December we look this way to see the night sky Figure 16.3 How our view of the night sky changes during the year The Sun and the Moon are the biggest and brightest objects in the sky. In comparison, all stars appear only as points of light. The closest planets may just appear as discs (rather than points) of light, especially Venus which is the brightest natural object in the night sky (other than the Moon). There are a few other things we might see in the night sky. At certain times, if we are lucky, we may also be able to see a comet, an artificial satellite or a meteor – which causes the streak of light seen in the sky when a rock fragment enters the Earth’s atmosphere and burns up due to friction. Occasionally, parts of meteors are not completely vaporized and they reach the Earth’s surface. They are then called meteorites and are extremely valuable for scientific research, being a source of extra-terrestrial material. 16.1 (D1: Core) Stellar quantities  Astrophysics Internet sites For many students astrophysics is a fascinating subject, but the opportunities for practical work are obviously limited. However, a considerable amount of very interesting information and stunning images are available on the internet and, without doubt, it will greatly enhance the study of this topic if you have easy and frequent access to the websites of prominent space organizations such as the European Space Agency (ESA), NASA and Hubble among others. Objects in the universe In this option we will concentrate our attention on (in order of size): planets and planetary systems (like the solar system), including comets stars (single and binary), stellar clusters (open and globular) and constellations nebulae galaxies, clusters of galaxies and super clusters of galaxies. Nebulae Nebulae are enormous diffuse ‘clouds’ of interstellar matter, mainly gases (mostly hydrogen and helium) and dust. Some of the matter may be ionized. A nebula forms over a very long time because of the gravitational attraction between the masses involved. (‘Interstellar’ means between the stars.) There are several kinds of nebulae, with different origins and different sizes. Large nebulae are the principal location for the formation of stars and most nebulae already contain stars that are the source of the energy and light by which they can be observed. It is possible to see some nebulae in our galaxy without a telescope, although they are diffuse and dim. They were probably first observed nearly 2000 years ago. Recent images of nebulae taken from the Hubble telescopes are truly spectacular. Figure 16.4 shows a telescope image of the Orion nebula. This can be seen without a telescope (close to Orion’s belt in the Orion constellation) and it contains a number of ‘young’ stars. It is one of the closest nebulae to Earth and one of the brightest, so it has been much studied as a source of information about the formation of stars. It is about 1 × 1016 km from Earth and about 2 × 1014 km in diameter, so that it subtends an angle at the eye of approximately 0.02 rad (≈ 1°, which is large in terms of astronomy). Figure 16.4 The Orion nebula (as pictured through a telescope) 3 4 16 Astrophysics Stars Within part of a nebula, over a very long period of time, gravity pulls atoms closer together and they can gain very high kinetic energies (that is, the temperature is extremely high – millions of kelvin) if the overall mass is large. The hydrogen nuclei (protons) can then have enough kinetic energy to overcome the very high electric forces of repulsion between them and fuse together to make helium nuclei. This process, known as nuclear fusion, can be simplified to: 411 H → 42 He + 201 e + neutrinos and photons Nuclear fusion happens in all stars (until near the end of their ‘lifetimes’) and is their dominant energy transformation. Each completed nuclear fusion of helium from four hydrogen nuclei (protons) is accompanied by a decrease in mass and an equivalent release of energy amounting to about 27 MeV (Chapter 12). The fusion of heavier elements occurs later in the lifetime of stars. When nuclear fusion begins on a large scale it is commonly described as the birth of a star. The thermal gas pressure (and radiation pressure) contraction of the material in the forming star creates outwards a thermal gas pressure and the emitted radiation also creates a radiation pressure outwards in opposition gravitational pressure to the gravitational pressure inwards. These pressures inwards remain equal and opposite for a very long time, during which the star will remain the same size, stable and unchanging. It will be in stellar equilibrium (Figure 16.5). It may be helpful to compare this to a balloon in equilibrium under the action of the gas pressure outwards and the pull of the elastic inwards. There is also a balance Figure 16.5 A stable star in equilibrium between energy transferred from fusions and energy radiated from the surface. During this period the star is known as a main sequence star. The only fundamental difference between these stars is their masses. Eventually the supply of hydrogen will be used up and the star will no longer be in equilibrium. This will be the beginning of the end of the ‘lifetime’ of a main sequence star. What happens then depends on the mass of the star (explained later in this chapter). Our Sun is approximately halfway through its lifetime as a main sequence star. Binary stars It is estimated that around half of all stars are in fact two (or more) stars orbiting around their common centre of mass with a constant period. Stars in a two-star system are described as binary stars (see Figure 16.6). Binary stars that are not too far away from Earth may be seen through a telescope as two separate Figure 16.6 An artist’s impression of a visual binary stars, but most binary stars are further away and appear as a single star system point of light. Binary star systems are important in astronomy because the period of their orbital motion is directly related to their mass. This means that if we can measure their period, we can calculate their mass. For non-visual binaries this may be possible using one of two observations: If one star passes regularly in front of the other as seen from Earth (an eclipse), the brightness will change periodically. If one star is momentarily moving towards the Earth, the other must be moving in the opposite direction. The frequency of the light received on Earth from each will be Dopplershifted (Chapter 9) periodically. 16.1 (D1: Core) Stellar quantities  Groups of stars Galaxies When we look at the stars in the night sky, they seem to be distributed almost randomly, but we are only looking at a tiny part of an enormous universe. The force of gravity causes billions of stars to collect into groups, all orbiting the same centre of mass. These groups are known as galaxies. Some of the spots of light we see in the night sky are distant galaxies (rather than individual stars). Billions of galaxies have been observed using astronomical telescopes. The Earth, the Sun and all the other stars that we can see with the unaided eye are in a galaxy called the Milky Way. Galaxies are commonly described by their shape as being spiral (Figure 16.7), elliptical or irregular. Galaxies are distributed throughout space, but not in a completely random way. For example, the Milky Way is one of a group of about 50 galaxies known as the ‘Local group’. Larger groups of galaxies, called clusters of galaxies, are bound together by gravitational forces. (See Figure 16.8 for an example.) Clusters may contain thousands of galaxies and much intergalactic gas along with undetected ‘dark matter’. (The term ‘galactic cluster’ is commonly used for a group of stars within a galaxy.) Clusters of galaxies are not distributed evenly throughout space, but are themselves grouped together in what are known as super clusters. Super clusters of galaxies may be the largest ‘structures’ in the universe. Figure 16.7 Spiral galaxy M81 Stellar clusters Some stars within a galaxy are close enough to each other that they become gravitationally bound together and rather than move independently, they move as a group called a stellar cluster. All the stars within a particular cluster were formed from the same nebula. There are two principal types of stellar cluster:  Globular clusters are old and contain many thousands of stars in roughly spherical shapes that are typically about 1014 km in diameter.  Open clusters are not as old as globular clusters. They are about the same size but contain much fewer stars (typically a few hundred). Because there are fewer stars in an open cluster, the overall shape is less well defined and the gravitational forces are weaker. Over time, an open stellar cluster may disperse because of the effects of other gravitational forces. The Pleiades (Figure 16.9) are an open cluster that is visible from Earth without a telescope. Figure 16.8 Virgo cluster of galaxies 5 6 16 Astrophysics Figure 16.9 The Pleiades are an open stellar cluster. It is important not to confuse stellar clusters, which are groups of stars relatively close to each other in space, with constellations. Constellations Ancient societies, such as Chinese, Indian and Greek civilizations, attempted to see some order in the apparent random scattering of the stars that we can see from Earth. They identified different parts of the night sky by distinguishing patterns of stars representing some aspect of their culture, such as the Greek hunter Orion (see Figure 16.10). Figure 16.10 The constellation of Orion: (a) the stars seen in the sky, (b) a representation from mythology These two-dimensional patterns of visible stars are called constellations. It is important to understand that the stars within any given constellation do not necessarily have anything in common. They may not even be ‘close together’, despite the impression we have by viewing them from Earth. Although many constellations were first named thousands of years ago, their names are still widely used today to identify parts of the night sky. 16.1 (D1: Core) Stellar quantities  Planetary systems around stars A planetary system is a collection of (non-stellar) masses orbiting a star. Planetary systems are believed to be formed by the same processes as the formation of the stars. Planets are objects of sufficient mass that gravitational forces have formed them into spherical shapes, but their mass is not large enough for nuclear fusion to occur. In other words, they are not massive enough to be stars. To distinguish planets from some smaller orbiting masses, it has been necessary for astronomers to specify that a planet has ‘cleared its neighbourhood’ of smaller masses close to its orbit. The search for extraterrestrial intelligence (SETI) concentrates on planetary systems like our own solar system and new planetary systems are now discovered regularly. By the start of the year 2014 more than one thousand planetary systems had been identified. In April 2014 astronomers announced that they had discovered the ‘most Earth-like planet’, Kepler 186f, orbiting a small star at a distance of about 500 ly (light years) from Earth (see Figure 16.11). Figure 16.11 An artist’s impression of Kepler 186f The solar system The Sun and all the objects orbiting it are collectively known as the solar system. Our Sun is a star and it is very similar to billions of other stars in the universe. It has many objects orbiting around it that are held in their orbits by gravity. The solar system is an example of a planetary system. Most of the planets have one or more objects orbiting around them. These are called moons. The Sun is the only large-scale object in our solar system which emits visible light; the others are only visible because they reflect the Sun’s radiation towards Earth. The Sun was formed about 4.6 billion years ago from the collapse of an enormous cloud of gas and dust. Evidence from radioisotopes in the Earth’s surface suggests that the Earth was formed about the same time, 4.5 billion years ago. Table 16.1 shows some details of the planets of our solar system (which do not need to be remembered). The distances given in the table are only averages because the planets are not perfect spheres and because their orbits are elliptical (oval) rather than circular. The Earth’s orbit, however, is very close to being circular so the Earth is always about the same distance from the Sun. (The Earth is closest to the Sun in January but there is only about a 3% difference between the smallest and largest separations.) An ellipse has two focuses (foci) and the Sun is located at one of those two points. The period of the Earth’s orbit is, of course, one year, but note that the further a planet is from the Sun, the longer its period. The link between orbital radius and period was discussed in Chapters 6 and 10. 7 8 16 Astrophysics Table 16.1 Planetary data (all data is correct to two significant figures) Mass/1024 kg Radius of planet/10 6 m Mean distance from Sun/1011 m Mercury 0.33 2.4 0.58 0.24 Venus 4.9 6.1 1.1 0.62 Planet Period/y Earth 6.0 6.4 1.5 1.0 Mars 0.64 3.4 2.3 1.9 Jupiter 1900 69 7.8 12 Saturn 570 57 14 29 Uranus 87 25 29 84 Neptune 100 25 45 160 Compared with planets, comets are relatively small lumps of rock and ice that also orbit the Sun, but typically with very long periods and very elliptical paths (see Figure 16.12). They therefore spend relatively little of their time in the inner solar system close to the Sun and the inner planets, such as Earth. When they approach the Sun, radiation and the outflow of particles (solar wind) often cause a comet to develop a diffuse tail of dust and gas, which always points away from the Sun (Figure 16.13). This, together with the rarity of seeing them, has made comets a matter of great curiosity for many of the world’s civilizations. Probably the most famous comet is named after the British astronomer and mathematician Edmund Halley (1656–1742). Halley correctly predicted that this comet would next be seen in 1758 (which was 16 years after his death). Halley’s comet has a period of 75 years; it was last seen in 1986 and will be seen next in the year 2061. In November 2014, after a 10-year mission, the European Space Agency’s spacecraft Rosetta landed the first object on a comet. The Philae lander was able to identify organic molecules on comet 67P. outer planet comet Sun Earth (not to scale) Figure 16.12 The eccentric (‘flattened’) path of a comet Figure 16.13 A comet and its tail 1 a Calculate the average density of Earth and Jupiter. b Why are they so different? 2 a What is the average orbital speed of the Earth? b Compare the Earth’s speed to that of Mercury. 3 a If there was a planet located at 35 × 1011 m from the Sun, suggest how long it might take to complete its orbit. b Would such a planet be visible to the unaided eye? Explain your answer. 4 a What is the smallest planet and what is its mass? b Why is Pluto not considered to be a planet? 5 What is the largest planet and what is its diameter? 16.1 (D1: Core) Stellar quantities  Additional Perspectives Asteroids colliding with the Earth Asteroids are large rocks that are generally bigger than comets but much smaller than planets. They do not have ‘tails’ and most orbit the Sun in approximately circular orbits between Mars and Jupiter, in a zone called the asteroid belt. Because they are relatively small, the trajectories (paths) of asteroids and comets may be significantly altered if they pass ‘close’ to a planet (especially Jupiter) when they are subject to large gravitational forces. Science-fiction authors and movie makers enjoy frightening us all with stories about asteroids or comets colliding with the Earth, but it is only in recent years that scientists have come to realise that such a major collision is not as unlikely as they had previously thought. In 1994 a large comet (Shoemaker– Levy 9) collided with Jupiter. The effect of the impact was seen easily with a telescope and was broadcast around the world on television (Figure 16.14). If a similar comet collided with Earth, the results would be catastrophic, although not quite on a scale comparable to the asteroid collision with Earth about 65 million years ago, which is thought to have led to the extinction of many species, including the dinosaurs. We only need to look at the cratercovered surface of the Moon Figure 16.14 Astronomers watch the impact of comet Shoemaker–Levy 9 with Jupiter. to become aware of the effects of collisions with asteroids and comets, but similar evidence is not so easy to find on the Earth’s surface. Rocks of diameter 10 m or less usually break up in the Earth’s atmosphere before impacting, so an asteroid would need to have a diameter of about 50 m or more before its impact would leave a noticeable and long-lasting crater. The effects of friction with the air might also cause an asteroid to explode before it impacted the Earth’s surface. Of course, most of the Earth is covered with water and no craters would be formed after an impact with the oceans. Also, old craters may well have been eroded, weathered or just covered with vegetation over long periods of time. Actual estimates about the size of possible asteroids that could collide with Earth and the probability of such events occurring are continually being refined. But, in general terms, we know that the probability of the Earth being struck by an asteroid is inversely related to its size. For example, an asteroid 50 m in diameter may impact the Earth about every 1000 years; a 1 km asteroid about every 500 000 years and a 10 km asteroid once every 100 000 000 years. The chance of a catastrophic impact in an average human lifetime may be about 1 in 10 000. There may be up to a million asteroids in our solar system capable of destroying civilization if they impacted with Earth, but it is not easy to observe all of them, nor track their movements. Much effort is now going into Near Earth Objects programs and researching what might be done if a dangerous impact was expected. 1 Calculate the kinetic energy of an asteroid of diameter 1 km and average density of 400 kg m–3 travelling at a speed of 20 km   s–1. Compare your answer with 25 megatonnes of TNT, the energy that would be released from a ‘large’ nuclear bomb. (1 tonne of TNT is equivalent to 4.2 × 109   J.) 2 Use the internet to find out when the next large asteroid is expected to pass near to Earth. How close will it come and how dangerous would it be if it hit us? 9 10 16 Astrophysics Astronomical distances The universe is enormous! Rather than use metres (or km) to measure distances, astronomers usually prefer to deal with smaller numbers and have introduced alternative units for distance. The light year, ly, is defined as the distance travelled by light in a vacuum in one year. At a light speed of 2.998 × 108 m s–1 and 365.25 days, a light year is easily shown to be 9.46 × 1015 m. This value is provided in the Physics data booklet. The astronomical unit, AU, is equivalent to the mean distance between the Earth and the Sun, 1.50 × 1011 m. This value is provided in the Physics data booklet. (Although the actual distance varies, the value of 1 AU is defined to be 1.495  978  707 × 1011 m.) One parsec, pc, is equal to 3.26 ly. This value is provided in the Physics data booklet. The parsec is the preferred unit of measurement in astronomy because it is closely related to parallax angles –the way in which the distances to ‘nearby’ stars are measured (this will be explained later). One parsec is defined as the distance to a star that has a parallax angle of one arc-second. While distances to ‘nearby’ stars are commonly measured in parsecs, the more distant stars in a galaxy are kpc away and distances to the most distant galaxies will be recorded in Mpc and Gpc. Table 16.2 Summary of distance units commonly used in astronomy Unit Metres/m 1011 Astronomical units/AU Light years/ly – – 1 AU = 1.50 × 1 ly = 9.46 × 1015 6.30 × 10 4 1016 105 1 pc = 3.09 × 2.06 × – 3.26 The scale of the universe The diameter of the observable universe is about 9 × 1010 ly. The speed of light limits the amount of the universe that we can, in principle, ‘observe’. The distance to the edge of the observable universe is equal to the speed of light multiplied by the age of the universe (but the expansion of space itself must be considered, which will be discussed later in the chapter). Distances between stars and between galaxies vary considerably. As a very approximate guide there might be 1012 stars in a big galaxy. A typical separation of stars within it may be about 1 ly, with a typical total diameter of a galaxy being about 104 ly (Figure 16.15). The billions of galaxies are separated from each other by vast distances, maybe 107 ly or more. the observable universe 9 × 1010 ly 106 ly 104 ly Typical stars may be 1 ly (or more) apart. There may be 1012 stars in a big galaxy Figure 16.15 Very approximate dimensions of galaxies (not to scale) 16.1 (D1: Core) Stellar quantities  11 6 Use Table 16.1 to determine the mean distance (in AU) from the Sun to the planets Mercury and Uranus. 7 What is the approximate size of the observable universe in: a km b pc? 8 Proxima Centauri is the nearest star to Earth at a distance of 4.0 × 1016 m. a How many light years is this? b If the Earth was scaled down from a diameter of 1.3 × 107 m to the size of a pin head (1 mm diameter), how far away would this star be on the same scale? 9 Our solar system has an approximate size of at least 1011 km. a How many light years is that? b If you were making a model of our solar system using a ball of diameter 10 cm to represent the Sun, how far away would the ‘edge’ of the solar system be? (The Sun’s diameter = 1.4 × 10 6 km.) c Research into how the edge of the solar system can be defined and what objects in the solar system are the most distant from the Sun. 10 Calculate the time for light to reach Earth from the Sun. 11 a Estimate how long would it take a spacecraft travelling away from Earth at an average speed of 4 km s–1 to reach: i Mars ii Proxima Centauri. b Find out the highest recorded speed of a spacecraft. 12 Use the data from Figure 16.15 to make a very rough estimate of the number of stars in the observable universe. 13 Research the diameter of our galaxy, the Milky Way, in parsecs. 14 Explain why it would be unusual to quote a distance between stars in AU. ToK Link Imagination The vast distances between stars and galaxies are difficult to comprehend or imagine. Are other ways of knowing more useful than imagination for gaining knowledge in astronomy? Imagining the vast distances in the universe may be considered to be similar to imagining the number of molecules in a grain of salt – the numbers are so large that they are almost meaningless to us. There is no doubt that it does help us to make comparisons like ‘it would take more than a billion years to walk to the nearest star’, but then we realise that this is an incredibly small distance in the universe! Determining the distances to the stars and distant galaxies The measurement of astronomical distances is a key issue in the study of astronomy. However, determining the distance from Earth to a star or galaxy accurately is not easy and a variety of methods have been developed. In this course we will consider three different ways in which the distance to a star or distant galaxy may be determined: stellar parallax use of Cepheid variable stars use of supernovae. The use of stellar parallax for ‘nearby’ stars is the most direct and easily understood method. The other two methods are used for distant galaxies. They will be discussed later in the chapter. Stellar parallax and its limitations This method is similar in principle to one that we might use on Earth to determine the distance to an inaccessible object, such as a boat or a plane. If the object can be observed from two different places, then its distance away can be calculated using trigonometry. An example of this triangulation method is shown in Figure 16.16. 16 Astrophysics An observer on land sees the boat from position P and then moves to position Q. If the angles α and β are measured and the distance PQ is known, then the other distances can be calculated. When astronomers want to locate a star, they can try to observe it from two different places, but the distance between two different locations on Earth is far too small compared with the distance between the Earth and the star. Therefore, astronomers observe the star from the same telescope at the same location, but at two different places in the Earth’s orbit; in other words, at different times of the sea α β year. To get the longest difference they usually take two measurements separated in time by six months. land P Q The triangulation method described above to locate a Figure 16.16 Determining the distance to a ship at sea boat would be much more difficult if the observer was in a using triangulation moving boat at sea and this is similar to the difficulty faced by astronomers locating stars from Earth. The problem can be overcome by comparing the position of the star to other stars much further away (in the ‘background’). This is known as a parallax method. This nearby star Parallax is the visual effect of a nearby object appearing seems to change fixed position during to move its position, as compared to more distant objects pattern the year (behind it), when viewed from different positions. A simple of stars example is easily observed by looking at one finger held in front of your face and the background behind it, first with one eye and then the other. In the same way, a ‘nearby’ star can appear to very slightly change its position during the year compared to other stars much further away (although, as we Figure 16.17 A nearby star’s apparent movement due to have noted before, stars generally appear to remain in fixed parallax patterns over very long periods of times). Stellar parallax (Figure 16.17) was first confirmed in 1838. Many astronomers had tried to detect it before (without success) because the existence of stellar parallax provides evidence for the motion of the Earth around the Sun. Using telescopes, astronomers measure the parallax angle, p, between, for example, observations of the star made in December and June. Figure 16.18 shows the angular positions of a nearby star in December and June. (In Figures 16.18 and 16.19 the size of the parallax angle has been much exaggerated for the sake of clarity.) Figure 16.18 Measuring the parallax angle six months apart nearby star position in June position in December p p parallax angle parallax angle Earth in December telescope p p distance to star, d 12 Sun 1 AU 1 AU Earth in June Figure 16.19 The geometry of the parallax angle 16.1 (D1: Core) Stellar quantities  13 If the measurements are made exactly six months apart, the distance between the locations where the two measurements are taken is the diameter of the Earth’s orbit around the Sun. We may assume that the orbit is circular, so that the radius is constant. The parallax of even the closest stars is very small because of the long distances involved and this means that the parallax angles are so tiny that they are measured in arc-seconds. (There are 3600 arc-seconds in a degree.) Once the parallax angle has been measured, simple geometry can be used to calculate the distance to the star (Figure 16.19): 1.50 × 1011 (m) V. I. relation d Note that the distance from the Earth to the star and the distance from the Sun to the star can be considered to be equal for such very small angles, so: p (rad) = sin p = tan p. parallax angle, p (rad) = Worked example 1 Calculate the distance, d, to a star if its parallax angle is 0.240 arc-seconds. π = 1.16 × 10 –6 rad 0.240 arc-seconds = 0.240 × 3260 36 0 0 ( ) ( ) 1011 1.50 × d (m) 1.50 × 1011 1.16 × 10 –6 = d d = 1.29 × 1017 m (= 13.7 ly) p (rad) = V. I. Example... If a parallax angle can be measured for a nearby star, calculations like this can be used to determine its distance away. Such calculations are common and it is much easier to use the angle directly as a measure of distance rather than making calculations in SI units. We have already seen that the parsec is defined as the distance to a star that has a parallax angle of one arc-second. But there is an inverse relationship here – larger parallax angles mean smaller distances. So: d (parsec) = 1 p (arc-second) This equation is given in the Physics data booklet. For example, a star with a parallax angle, p, 1 = 4 pc away etc. Table 16.3 shows the of 0.25 arc-seconds will be from a star which is 0.25 relationship between parallax angle and distance. The stellar parallax method is limited by the inability of telescopes on Earth to observe very small shifts in apparent positions of stars or accurately measure very small angles less than 0.01 arc-seconds. Table 16.3 Parallax angles in arc-seconds and distances in parsecs This means that this method is usually limited to those stars that are relatively close to Earth, within Parallax angle /arc-seconds Distance away/pc about 100 pc, well within our own galaxy. The use 0.10 10.00 of telescopes on satellites above the turbulence and 0.25 4.00 distortions of the Earth’s atmosphere can extend the 0.50 2.00 range considerably, but it is still not suitable for the 1.00 1.00 majority of stars, which are much further away. 14 16 Astrophysics 15 Convert an angle of 1 arc-second to: a degrees b radians. 16 The parallax angle for Barnard’s star is measured to be 0.55 arc-seconds. How far away is it from Earth: a in pc b in m c in ly? 17 What are the parallax angles for three stars at the following distances from Earth? a 2.47 × 1015 km b 7.9 ly c 2.67 pc Luminosity and apparent brightness Every star (apart from our Sun) appears to us as a point in space. The only direct information that we can have about any particular star is its position (as might be displayed on a twodimensional star map), the intensity of radiation received from it and the spectrum of its radiation. These are the only observable differences between all the stars that we can detect. The apparent brightness, b, of a star (including the Sun) is defined as the intensity (power/area) received (perpendicular to direction of propagation) at the Earth. The units are W m–2. The apparent brightness of the Sun is approximately 1360 W m–2 above the Earth’s atmosphere. This is also called the solar constant and was discussed in Chapter 8. Of course, the apparent brightnesses of all the other stars are much, much less. A typical value would be 10 – 12 W m–2. Astronomers have developed very accurate means of measuring apparent brightnesses using charge-coupled devices (CCDs), in which the charge produced in a semiconductor is proportional to the number of photons received, and hence the apparent brightness. In Figure 16.20, stars A and B appear to be close together but in reality, in three-dimensional space, star A could be much closer to star C than star B. The situation may be further confused by differences in the brightness of the three stars. For example, it is feasible that star B could be the furthest away of these three stars and only appears brightest because it emits much more light than the other two. A B C Figure 16.20 The apparent brightnesses of three stars (as indicated by the diameters of the dots) The luminosity, L, of a star is defined as the total power it radiates (in the form of electromagnetic waves). It is measured in watts, W. surface area of sphere = 4πd2 star of luminosity, L apparent brightness, b = d Figure 16.21 Relating apparent brightness to luminosity L 4πd2 For example, the luminosity of the Sun is 3.8 × 1026 W. The apparent brightness of a star as observed on Earth will depend on its luminosity and its distance from Earth. We would reasonably expect that the energy from any star spreads out equally in all directions, so the power arriving at a distant observer on Earth will be very considerably less than the power emitted. Assuming that none of the emitted energy is absorbed or scattered as it travels across space, the power received per square metre anywhere on a sphere of radius d will be equal to the emitted power (luminosity) divided by the ‘surface’ area of the sphere, as shown in Figure 16.21. 16.1 (D1: Core) Stellar quantities apparent brightness, b = L 4πd2 V. I. Law This important equation is given in the Physics data booklet. Nature of Science The inverse square relationship The importance of this equation for apparent brightness lies in the fact that once we have measured the apparent brightness of a star and if we know its distance from Earth, then it is a simple matter to calculate the luminosity of the star. Conversely, as we shall see later, if the luminosity of a star is known, measurement of its apparent brightness can lead to an estimate of its distance from Earth. The information provided by this simple equation is fundamental to an understanding of basic astronomy. This is example of an inverse square relationship. If the distance from a star is multiplied by 2, then the apparent brightness is divided by 22; if the distance is multiplied by, for example, 37, then the apparent brightness will be divided by 372 (= 1369) etc. This is illustrated in Figure 16.22, which shows that at three times the distance, the same power is spread over nine (32) times the area. star intensity, I intensity, 14 I intensity, 19 I d 2d 3d Figure 16.22 How intensity changes with the inverse square law Not surprisingly, very little radiation is absorbed or scattered as it travels billions of kilometres through almost empty space, although the effects of the journey must be considered when studying the most distant galaxies. However, 100 km of the Earth’s atmosphere does have a very significant effect, reducing brightness and resolution in many parts of the spectrum. That is why astronomers often prefer to use telescopes sited on mountain tops or on satellites above the Earth’s atmosphere to gather data. Utilizations Telescopes on the ground and telescopes in orbit Waves from all parts of the electromagnetic spectrum arrive at the Earth from outer space and it is truly impressive to consider just how much scientists have learned about the universe from studying these various radiations. Most of this option is about how that information is interpreted, but little has been included about how waves from the various parts of the electromagnetic spectrum provide different information about their sources. Figure 16.23 shows a telescope designed to focus and detect radio waves from outer space. When radiation passes through the Earth’s atmosphere some of it may be absorbed, refracted or scattered, and these effects will often depend on the wavelengths involved. For example, 15 16 Astrophysics in visible light, the blue end of the spectrum is scattered more than red light and that helps to explain blue skies and red sunsets. We only have to look through the shifting haze above a hot surface to appreciate just how much convection currents in the air affect what we see. Figure 16.23 A telescope at the Very Large Array, New Mexico, USA receiving radio waves from space Astronomers have long understood the advantages of siting optical telescopes on the tops of mountains to reduce the adverse effects of the atmosphere on the images seen (Figure 16.24). The highest mountains are, of course, much lower than the height of the atmosphere, which is usually assumed to be approximately 100 km, although there is no distinct ‘edge’. The use of telescopes on orbiting satellites has greatly increased the resolution of images from space. (The resolution of images was discussed in detail in Chapter 9 and is needed by Higher Level students only.) The Hubble telescope (Figure 16.25) has been the focus of much attention, with many of its spectacular images well known around the world. The telescope was launched in 1990 and was named after the famous American astronomer, Edwin Hubble. It has a mass of about 11 tonnes and orbits approximately 560 km above the Earth’s surface, taking 96 minutes for one complete orbit. One of the finest achievements of astronomers using the Hubble telescope has been determining the distances to very distant stars accurately, enabling a much improved estimate for the age of the universe. The second major advantage of placing a satellite in orbit is that it can detect radiations that would otherwise be absorbed in the atmosphere before reaching any terrestrial telescopes (those on the Earth’s surface). Figure 16.26 indicates (approximately) the effect that the Earth’s atmosphere has on preventing radiations of different wavelengths from reaching the Earth’s surface. 1 M  ake a sketch of Figure 16.26 and indicate and name the different sections of the electromagnetic spectrum. Figure 16.24 The telescopes at the Paranal Observatory on the top of Cerro Paranal, a mountain in the Atacama Desert in Chile 2 V  isit the Hubble website to look at the magnificent images from space and make a list of the important characteristics of the telescope. Percentage of radiation reaching sea level 16 100 50 0 10–10 Figure 16.25 The Hubble telescope 10–8 10–6 10–4 10–2 1 102 Wavelength, λ/m Figure 16.26 How the Earth’s atmosphere affects incoming radiation 16.1 (D1: Core) Stellar quantities  17 Worked example 2 A star of luminosity 6.3 × 1027 W is 7.9 × 1013 km from Earth. What is its apparent brightness? L b= 4πd2 6.3 × 1027 Good example !!!! b= 4π × (7.9 × 1016)2 b = 8.0 × 10 –8 W m –2 18 How far away from Earth is a star that has a luminosity of 2.1 × 1028 W and an apparent brightness of 1.4 × 10 –8 W m –2? 19 A star that is 12.4 ly from Earth has an apparent brightness of 2.2 × 10 –8 W m –2. What is its luminosity? 20 Calculate the distance to the Sun using values for its luminosity and apparent brightness. 21 Star A is 14 ly away from Earth and star B is 70 ly away. If the apparent brightness of A is 3200 times higher than that of star B, calculate the ratio of their luminosities. 22 If the radiation from the star in Question 18 has an average visible wavelength of 5.5 × 10 −7 m, estimate how many visible photons arrive every second at a human eye of pupil diameter 0.50 cm. Relating the luminosity of a star to its surface temperature We know from Chapter 8 that the power radiated from a surface in the form of electromagnetic waves can be calculated from P = eσAT4 where A is the surface area, T is the temperature (K) and σ is the Stefan Boltzmann constant. We may assume that stars behave as perfect black bodies, so that emissivity e = 1 and the power emitted by a star (its luminosity, L) is then given by: L = σAT 4 This equation is given in the Physics data booklet. Remember that when a surface is described as a ‘perfect black body’ we mean that it emits the maximum possible radiation at any particular temperature and not that it appears black. This equation shows us that if we know the luminosity of a star and its surface temperature then we can calculate its surface area and radius (A = 4πr2). This is shown in the following worked example and questions. In the next section we will review how Wien’s displacement law (Chapter 8) can be used to determine the surface temperature of a star from its spectrum. Worked example V. I. Example 3 What is the luminosity of a star of radius 2.70 × 106 km and surface temperature 7120 K? L = σAT4 = (5.67 × 10 −8) × 4π × (2.70 × 106 × 103)2 × (7120) 4 = 1.33 × 1028 W 23 A star has a surface area of 1.8 × 1019 m2 and a surface temperature of 4200 K. What is its luminosity? 24 If a star has a luminosity of 2.4 × 1028 W and a surface temperature of 8500 K, what is: a its surface area b its radius? These are very important questions. Good for the students 25 What is the surface temperature of a star that has an area of 6.0 × 1020 m2 and a luminosity of 3.6 × 1030 W? 26 If the star in question 23 is 17.3 ly away, what will its apparent brightness be when seen from Earth? 27 If the star in question 24 has an apparent brightness of 2.5 × 10 −8 W m –2, how many kilometres is it from Earth? 28 Compare the luminosities of these two stars: star A has a surface temperature half that of star B, but its radius is forty times larger. 29 A star has eighty times the luminosity of our Sun and its surface temperature is twice that of the Sun. How much bigger is the star than our Sun? 16 Astrophysics 16.2 (D2: Core) Stellar characteristics and stellar evolution – a simple diagram that plots the luminosity versus the surface temperature of stars reveals unusually detailed patterns that help understand the inner workings of stars; stars follow well-defined patterns from the moment they are created to their eventual death Stellar spectra Nature of Science Figure 16.27 The black-body spectra emitted by stars with different surface temperatures Evidence provided by spectra Astronomers have learnt a great deal about the universe from a limited range of evidence received from sources that are enormous distances from Earth. Apart from the location and luminosity of stars, a surprisingly large amount of information can be determined from close examination of the spectrum produced by a star. If we can measure the wavelength at which the emitted radiation has its maximum intensity, we can calculate the surface temperature of a star. If we can observe the absorption spectrum produced by the outer layers of a star, we can determine its chemical composition. If we compare the absorption spectrum received from a star to the spectrum from the same element observed on Earth, we can use the Doppler shift to determine the velocity of the star (or galaxy); this provides evidence for the expansion of the universe. (See section 16.3.) Intensity 18 T = 6000 K visible spectrum Surface temperature Figure 16.27 shows that the spectra from stars with different surfaces temperatures differ, not only in overall intensity, but also in the spread of wavelengths emitted. This graph is similar to one previously seen in Chapter 8. Wien’s displacement law was also discussed in Chapter 8. It is an empirical law that represents how the wavelength at which the radiation intensity is highest becomes lower as the surface gets hotter: T = 5500 K T = 5000 K T = 4500 K λmaxT = 2.9 × 10 –3 m K T = 4000 K T = 3500 K 0 500 1500 2000 Wavelength, λ/nm 1000 This equation is given in the Physics data booklet. It was also given in Chapter 8 in a slightly different form. increasing frequency Worked example 4 Very good example !!! What is the surface temperature of a star that emits radiation with a peak of intensity at a wavelength of 1.04 × 10 –7 m? λmaxT = 2.9 × 10 –3 m K (1.04 × 10 −7)T = 2.90 × 10 –3 2.90 × 10 –3 T= 1.04 × 10 –7 = 27 900 K 16.2 (D2: Core) Stellar characteristics and stellar evolution  19 30 If the surface temperature of the Sun is 5700 K, at what wavelength is the emitted radiation maximized? In what part of the visible spectrum is this wavelength? 31 A star emits radiation that has its maximum intensity at a wavelength of 6.5 × 10 –7 m. a What is its surface temperature? b If it has a luminosity of 3.7 × 1029 W, what is the surface area of the star? c What is its radius? 32 a At what wavelength does a star with a surface temperature of 8200 K emit radiation with maximum intensity? b If this star has a radius of 1.8 × 106 km, what is its luminosity? c If it is 36 ly from Earth, what is its apparent brightness? 33 The star Canopus has a luminosity of 5.8 × 1030 W and a radius of 4.5 × 1010 m. Use this data to estimate the wavelength at which it emits the most radiation. 34 Sketch graphs comparing the emission spectra from the stars Betelgeuse (3600 K) and Alkaid (20 000 K). Utilizations The classification of stars by the colours they emit The surface temperatures of different stars may be as cool as a few thousand kelvin or as hot as 40 000+ K. Although Figure 16.27 only shows graphs for cooler stars, it should be clear that the range of visible colours present in the spectra from stars at different temperatures will be slightly different. For example, the light produced by a surface temperature of 4500 K has its highest intensity in the red end of the spectrum, whereas the light produced by 6000 K has more from the blue-violet end of the spectrum. Hotter stars are blue/white and cooler stars are yellow/red. To observers on Earth this will be noticed as slight differences in colour and this has long been the way in which astronomers group and classify different stars. In general, cooler stars are slightly redder and hotter stars are slightly bluer. Table 16.4 lists the eight spectral classes into which all visible stars are placed. Table 16.4 Spectral classes, temperatures and colours Spectral class Surface temperature/K Colour O 30 000–50 000 blue B 10 000–30 000 blue–white A 7500–10 000 white F 6000–7500 yellow–white G 5000–6000 yellow K 3500–5000 yellow–red (orange) M 2000–3500 red V. I. Table This apparently haphazard system of lettering stars according to their colour is an adaptation of an earlier alphabetical classification. A widely quoted mnemonic for remembering the order (from the hottest) is ‘Only Bad Astronomers Forget Generally Known Mnemonics’. 1 a What is the spectral class of our Sun? b We often refer to the light from our Sun as ‘white’. Discuss whether or not this is an accurate description. 2 Two common types of star are called red giants and white dwarfs. What spectral class would you expect them to be? 3 What is the spectral class and colour of the star Alkaid (referred to in question 34)? 16 Astrophysics Chemical composition As the continuous black body spectrum emitted from a star passes through its cooler outer layers, some wavelengths will be absorbed by the atoms present. When the radiation is detected on Earth, an absorption spectrum (discussed in Chapter 7) will be observed. Because we know that every chemical element has its own unique spectrum, this information can be used to identify the elements present in the outer layers of a star. The element helium is the second most common in the universe (after hydrogen), but it was not detected on Earth until 1882. Fourteen years earlier, however, it had been identified as a new element in the Sun from its spectrum (see Figure 16.28). Figure 16.29 shows a graphical representation of how a black-body spectrum emitted Figure 16.28 The absorption by the core of a star is modified by absorption of radiation in the outer layers. spectrum of helium Intensity 20 absorption lines Wavelength Figure 16.29 Graph of intensity against wavelength for a stellar absorption spectrum ToK Link The role of interpretation The information revealed through spectra needs a trained mind to interpret it. What is the role of interpretation in gaining knowledge in the natural sciences? How does this differ from the role of interpretation in the other areas of knowledge? Without detailed scientific knowledge and understanding, the observation of spectra would offer no obvious clues about the nature of stars. This is equally true of many other aspects of astronomy. Without scientific expertise, the information is of no use and may seem irrelevant, so that a non-expert would probably be unable to comment meaningfully. The same comments apply to advanced studies in other scientific disciplines. The Hertzsprung–Russell (HR) diagram The luminosity of a star depends on its surface temperature and its surface area (L = σAT4), so a star could be particularly luminous because it is hot or because it is big, or both. In the early twentieth century two scientists, Hertzsprung and Russell, separately plotted similar diagrams of luminosity against temperature in order to determine whether or not there was any pattern in the way that the stars were distributed. If there were no similarities in the composition of different stars, they could have many different combinations of temperature, size and luminosity. This would lead to them being randomly distributed on a luminosity–temperature diagram. But, more than about 90% of all stars are undergoing the same processes, fusing hydrogen into helium (as explained previously) and they are in a similar kind of equilibrium. These stars are called main sequence stars and their only essential difference is their mass. 16.2 (D2: Core) Stellar characteristics and stellar evolution  21 Mass–luminosity relation for main sequence stars Stars that are formed from higher masses will have stronger gravitational forces pulling them together. This will result in higher temperatures at their core and faster rates of nuclear fusion. More massive main sequence stars will have larger sizes, higher surface temperatures and brighter luminosities. The relationship between luminosity, L, and mass, M, for main sequence stars is described by the following equation, which is given in the Physics data booklet. This is an approximate, generalized relationship and it may not be precise for any given star. Luminosity L ∝ M3.5 cooler and brighter hotter and brighter S AS G SIN M A RE C IN cooler and dimmer hotter and dimmer Temperature Figure 16.30 Linking mass, temperature and luminosity for main sequence stars For example, if star A has twice the mass of star B, star A will have a luminosity approximately 23.5 times greater than star B (≈ ×11). This means that the rate of nuclear fusion in the more massive star will be much higher and it will have a much shorter lifetime than a less massive star. If the relationship between the mass and luminosity of a star is represented by this relationship, then we can be sure that it is a main sequence star. Figure 16.30 suggests how we might expect a luminosity–temperature diagram to appear for main sequence stars of different masses. Hertzsprung and Russell plotted data from a very large number of stars on luminosity–temperature diagrams, but the important HR diagram has two significant differences from Figure 16.30: 1 For historical reasons the temperature scale is reversed. 2 Because of the enormous differences in the luminosity of stars, the scale is logarithmic, rather than linear. (The temperature scale is also usually logarithmic.) Figure 16.31 shows a large number of individual stars plotted on a Hertzsprung–Russell (HR) diagram, with all luminosities compared to the luminosity of our Sun (LΘ). This figure also tries to give an impression of the colours of the stars. Luminosity/LO 106 R= 10 R= RO 100 supergiants R= RO 100 0R O 105 104 R= 1R O red giants instability strip (Cepheid variables) 103 R= 102 10 R= 0.1 0.0 RO O 1 10–1 main sequence 1R R= 0.0 01 Sun RO 10–2 10–3 R= 0.0 001 RO white dwarfs 10–4 40000 30000 20000 Figure 16.31 The Hertzsprung–Russell (HR) diagram 10000 Surface temperature/K 5000 2500 22 16 Astrophysics It should be apparent that the stars are not distributed at random in the HR diagram. The diagram can be used as a basis for classifying stars into certain types. As already explained, most stars (about 90%) can be located in the central band, which runs from top left to bottom right in Figure 16.31. These are the main sequence stars. The 10% of stars which are not in the main sequence are important and they will be discussed in the next section. In general, we can say that any stars located vertically above the main sequence must be larger (than main sequence stars) in order to have higher luminosity at the same temperature. By similar reasoning, any stars vertically below the main sequence must be smaller than main sequence stars of the same temperature. By considering L = σAT4 and A = 4πR2 (leading to L = σ4πR2T4), it is possible to draw the lines of constant radius on the HR diagram (as shown in Figure 16.31). Worked example 5 Use the HR diagram in Figure 16.31 to predict the surface temperature of a main sequence star that has ten times the radius of our Sun. The band of main sequence stars crosses the R = 10R line at about 30 000 K. 35 The radius of the Sun is 7 × 10 8 m and its surface temperature is 5800 K. Estimate the radius of a main sequence star that has a surface temperature five times that of our Sun. 36 The luminosities of two main sequence stars are in the ratio 10 : 1. What is the ratio of their masses? 37 a b c d e A star has a mass five times heavier than the mass of our Sun. Estimate its luminosity. What assumption did you make? Which star will have the longer lifetime? Use the HR diagram in Figure 16.31 to estimate the surface temperature of the star. Approximately how many times bigger is this star than our Sun? 38 Use the HR diagram to estimate the difference in diameter of a white dwarf star and a supergiant star if they both have the same surface temperature. Utilizations Using the HR diagram to estimate the distance to stars Using Wien’s law the surface temperature of a star can be determined from its spectrum and, assuming that it is a main sequence star, it is a relatively simple matter to use the HR diagram to estimate its luminosity, L, and hence calculate its distance, d, away from Earth using b = L/4πd2 and a measurement of its apparent brightness, b. This method assumes that the radiation, which has travelled vast distances from very distant stars, has not been altered in any way by the journey. For example, if any radiation is absorbed or scattered during the journey, the value of apparent brightness used in calculations will be less than it would have been without the absorption or scattering, leading to an over-estimate of the distance to the star. Because the exact position of the star on the HR diagram may not be known with accuracy and because of unknown amounts of scattering/absorption, there is a significant uncertainty in this method of determining stellar distances. In fact, the use of this method (misleadingly called spectroscopic parallax) is mostly confined to our galaxy. The majority of stars are obviously further away in other galaxies, so to determine the distances to those galaxies we need other methods. 1 Estimate the distance from Earth (in pc) of a main sequence star that has a surface temperature of 7500 K and an apparent brightness of 4.6 × 10−13 W m–2. 16.2 (D2: Core) Stellar characteristics and stellar evolution  23 Types of star that are not on the main sequence Cepheid variables Maximum luminosity/LO Relative luminosity/% The instability strip on the HR diagram contains a number of different kinds of pulsating stars. Such stars have moved off the main sequence and are oscillating under the effects of the competing gravitational pressure, and radiation and thermal pressures. The most important stars in the instability strip are known as Cepheid variables. In a Cepheid variable the outer 100 layers regularly expand and contract (typically by 30%) with surprisingly short 80 periods (in astronomical terms), resulting in very regular and precise variations in luminosity (see Figure 16.32) – a typical 60 period is a few weeks. If the surface temperature remains approximately 40 constant then the increasing luminosity is explained by the larger surface area 20 when the star expands. Although Cepheid variables are 0 not common stars, they are important 0 5 10 15 and their behaviour has been studied Time/days in great depth. From observations Figure 16.32 Variation in luminosity of a Cepheid variable star of those Cepheid variables that are relatively close to Earth, it is known that there is a precise relationship between the time period of their pulses of luminosity (and hence their received apparent brightness on Earth) and the peak value of that luminosity. This was first discovered by Henrietta Leavitt (Figure 16.33) in 1908. This relationship is called the period–luminosity relationship and it is commonly presented in graphical form, as shown in Figure 16.34; the longer the period, the higher the luminosity of the Cepheid variable. 10 000 1000 100 1 Figure 16.33 Henrietta Leavitt discovered the periodicity of Cepheid variables in 1908. 10 100 Period/days Figure 16.34 Period–luminosity relationship for a Cepheid variable Note the logarithmic nature of both the scales on the graph in Figure 16.34. This is necessary to include the enormous range of values involved. Using Cepheid variables to determine astronomical distances If the luminosity of a Cepheid variable can be determined from its period, then its distance from Earth, d, can be calculated if its apparent brightness, b, is measured. Again we can use the equation: L b= 4πd2 24 16 Astrophysics Inaccuracies in the data involved mean that these estimates of distance, especially to the furthest galaxies, are uncertain. This uncertainty is a significant problem when estimating the age of the universe. Astronomers often describe Cepheid variables as ‘standard candles’ because if their distance from Earth is determined, it can then be taken as a good indication of the distance of the whole galaxy from Earth (since that distance is very much longer than the distances between stars within the galaxy, see Figure 16.15). Worked example 6 A Cepheid variable in a distant galaxy is observed to vary in apparent brightness with a period of 8.0 days. If its maximum apparent brightness is 1.92 × 10 –9 W m –2, how far away is the galaxy? From a luminosity–period graph (similar to Figure 16.34), the maximum luminosity can be determined to be 2500 times the luminosity of the Sun. luminosity = 2500 × (3.8 × 1026 W) = 9.5 × 1029 W L b= Very good example!!! 4πd2 9.5 × 1029 1.92 × 10 –9 = 4πd2 d = 6.3 × 1018 m 39 a If a Cepheid variable has a period of 15 days, what is its approximate maximum luminosity? b If the star is 3.3 Mpc from Earth, what is the maximum observed apparent brightness? 40 A Cepheid variable is 15 kpc from Earth and is observed to have a maximum apparent brightness of 8.7 × 10 –13 W m –2. a Calculate the maximum luminosity of this star. b Use Figure 16.34 to estimate the time period of the variation in the star’s luminosity. 41 For very large distances astronomers may use supernovae (rather than Cepheids) as ‘standard candles’. Suggest a property of supernovae which might be necessary for this. What happens to a star when the supply of hydrogen is reduced? Over a long period of time, the amount of hydrogen in the core of a star gets significantly less, so that eventually the outwards pressure is reduced and becomes smaller than the inwards gravitational pressure. This occurs when the mass of the core is about 12% of the star’s total mass and there is still plenty of hydrogen remaining in the outer layers of the star. The star begins to contract and gravitational energy is again transferred to kinetic energy of the particles (the temperature of the core rises even higher than before – to 108 K and higher). This forces the outer layers of the star to expand and consequently cool. At the higher temperatures in the core (in all but the very smallest stars) it is now possible for helium to fuse together to form carbon and possibly some larger nuclei, releasing more energy so that the star becomes more luminous. So, the star now has a hotter core but it has become larger and cooler on the surface. Its colour changes and it is now known as a red giant (or, if it is very large, a red supergiant). At this point it will leave the main sequence part of the HR diagram. All main sequence stars follow predictable patterns but the heavier the mass of a star, the higher the gravitational potential energy and the higher the temperature when it begins to collapse. We can identify three different outcomes, depending only on the mass of the original star: white dwarf, neutron star or black hole. Red giants, white dwarfs, neutron stars and black holes Red giants As explained above, most stars will become red giants (or red supergiants) at the end of their time on the main sequence. They are described as giant stars because they have expanded 16.2 (D2: Core) Stellar characteristics and stellar evolution  25 considerably from their original size and, in doing so, their surfaces have cooled and therefore changed in colour to slightly red. White dwarfs After nuclear fusion in the core finishes, if the mass of a red giant star is less than a certain limit (about eight solar masses), the energy released as the core contracts forces the outer layers of the star to be ejected in what is known as a planetary nebula. (Be careful with this name – it is misleading because it has nothing to do with planets.) The core of the star that is left behind has a much reduced mass and is described as a dwarf star. A process known as electron degeneracy pressure (electrons acting like a gas) prevents the star collapsing further so this kind of star can remain stable for a long time. Such stars are known as white dwarfs because they have low luminosities (they cannot be seen without a telescope), but their surface temperatures are relatively high (L = σAT4). Studying the patterns we see in other stars helps us to understand our own Sun and what will happen to it in the future. It is about halfway through its time as a main sequence star and it will become a red giant in about seven billion years, after which it will become a white dwarf. Neutron stars or black holes Red giants with original masses greater than about eight solar masses are known as red supergiants and they do not evolve into white dwarfs. The electron degeneracy pressure is insufficient to resist the gravitational forces and the gravitational potential energy released is so high that there are dramatic changes in the core that result in an enormous explosion called a supernova. Here again, the result depends on the mass involved. If the original mass of the star was between 8 and 20 solar masses, the remaining core after the supernova will form a neutron star. If the mass was greater, a black hole is formed. Neutron stars After a supernova of a red supergiant, if the remaining core has a mass between approximately 1.4 and 3 solar masses it will contract to a neutron star. Neutron stars are extremely dense (ρ ≈ 5 × 1017 kg m–3), but resist further compression due to a process called neutron degeneracy pressure. Black holes If the remnant after a supernova has a remaining mass of more than approximately three solar masses, neutron degeneracy pressure is insufficient to resist further collapse. The result is a black hole, which produces such strong gravitational forces that not even the fastest particles, photons (for example, light) can escape. Black holes cannot be observed directly, but they can be detected by their interaction with other matter and radiation. For example X-rays are produced when superheated matter spirals towards a black hole. NASA’s Chandra Observatory was designed to search for black holes. The first black hole was confirmed in 1971. Astronomers believe that our own galaxy, the Milky Way, has a supermassive black hole near its centre. Chandrasekhar and Oppenheimer–Volkoff limits The mass limits mentioned above are known by the names of prominent astronomers: The Chandrasekhar limit is the maximum mass of a white dwarf star (= 1.4 × solar mass). The Oppenheimer–Volkoff limit is the maximum mass of a neutron star (≈ 3 × solar mass). Figure 16.35 represents these limits in a simplified chart of stellar evolution. 16 Astrophysics Figure 16.35 Evolution of stars of different masses (the numbers shown represent the approximate mass limits of the stars as multiples of the current mass of the Sun) supernova Oppenheimer–Volkoff 20 limit black holes 3.0 red supergiants Increasing mass supernova main sequence of stars according to their masses Chandrasekhar limit neutron stars 1.4 8 red giants ejection of planetary nebula white dwarfs collapse of core hydrogen in core becomes depleted Stellar evolution on HR diagrams When a main sequence star expands to a red giant, or a red supergiant, its luminosity and surface temperature change and this, and subsequent changes over time, can be tracked on an HR diagram. It is known as a star’s evolutionary path. Typical evolutionary paths of low-mass and high-mass stars are shown in Figure 16.36. Figure 16.36 Evolutionary paths of stars after they leave the main sequence Luminosity 26 supernova red supergiant red giant ma in se q ue nc e large star small star white dwarf Temperature 42 Explain why neutron stars and black holes cannot be placed on a HR diagram. 43 Use the internet to learn more about electron and neutron degeneracy. 44 Suggest how a black hole can be detected, even though it cannot be seen. 45 Explain why the Chandrasekhar limit is such an important number in astronomy. 46 Explain why some supernovae result in neutron stars, while others result in black holes. 16.3 (D3: Core) Cosmology  27 16.3 (D3: Core) Cosmology – the Hot Big Bang model is a theory that describes the origin and expansion of the universe and is supported by extensive experimental evidence Cosmology is the study of the universe – how it began, how it developed and what will happen to it in the future. It has always been the nature of many individuals, societies and cultures to wonder what lies beyond the Earth. The fact that the Sun and the stars appear to revolve around the Earth led early civilizations, understandably but wrongly, to believe that a stationary Earth was the centre of everything. This belief was often fundamental to their religions. Indeed, even today some people still believe from their everyday observations, or their religious beliefs, that the Sun orbits around the Earth rather than the other way around. Nature of Science Additional Perspectives Models of the universe In the Newtonian model of the universe, the Earth, the Sun and the planets were just tiny parts of an infinitely large and unchanging (static) universe that had always been the way it is, and always would be the same. In this model, the universe, on the large scale, is more or less the same everywhere. In other words it is uniform with stars approximately evenly distributed. Newton reasoned that unless all of these assumptions (sometimes called postulates) were valid, then gravitational forces would be unbalanced, resulting in the movement of stars (which were thought to be stationary at that time). But there is a big problem with this Newtonian model of the universe, one that many astronomers soon realized. If the universe is infinite and contains an infinite number of stars, there should be no such thing as a dark sky at night, because light from the stars should be arriving from all directions at all times. (This is known as Olbers’s paradox, named after one of the leading astronomers of the nineteenth century, Heinrich Wilhelm Olbers. A paradox is an apparently true statement that seems to contradict itself. ‘I always lie’ is a widely quoted paradoxical statement.) It was clear that either the reasoning given above and/or the Newtonian model of the universe needed changing or rejecting. Since the mid to late 1960s, the Big Bang model of the universe has been widely accepted by astronomers and has solved Olbers’s paradox. ‘The shoulders of giants’ Nicolas Copernicus, a Polish astronomer and cleric (Figure 16.37), is considered by many to be the founder of modern astronomy. In 1530 he published a famous paper stating that the Sun was the centre of the universe and that the Earth, stars and planets orbited around it (a heliocentric model). At that time, and for many years afterwards, these views directly challenged ‘scientific’, philosophical and religious beliefs. It was then generally believed that the Earth was at the centre of everything (a geocentric model). That profound and widespread belief dated all the way back to Ptolemy, Aristotle and others nearly 2000 years earlier. It should be noted, however, that Aristrachus in Ancient Greece is generally credited with being the first well-known person to propose a heliocentric model. More than 100 years after the birth of Copernicus, and still before the invention of the telescope, an eccentric Danish nobleman, Tycho Brahe, became famous for the vast number of very accurate observations he made on the motions of the five visible planets. He worked mostly at an elaborate observatory on an island in his own country, but went to Prague a few years before his death in 1601. Johannes Kepler was Brahe’s assistant and later, after his death, he worked on Brahe’s considerable, but unexplained, data to produce his three famous laws of planetary motion. Figure 16.37 Copernicus At about the same time in Italy, the astronomer Giordano Bruno had taken the heliocentric model further with revolutionary suggestions 28 16 Astrophysics that the universe was infinite and that the Sun was not at the centre. The Sun was, Bruno suggested, similar in nature to the other stars. He was burned at the stake in 1600 for these beliefs – killed for his, so-called, heresy. About 30 years later, one of the greatest scientific thinkers of all time, Galileo Galilei, was placed on trial by the Roman Catholic Church under similar charges. Many years earlier he had used the newly invented telescope to observe the moons of Jupiter and had reasoned that the Earth orbited the Sun in a similar way, as had been proposed by Copernicus. Under pressure, he publicly renounced these beliefs and was allowed to live the rest of his life under house arrest. All this has provided the subject of many books, plays and movies. 1450 1475 1500 1525 1550 1575 1600 Copernicus Brahe (1473–1543) (1546–1601) Bruno 1625 1650 1675 1700 1725 1750 (1548–1600) Kepler (1571–1630) Galileo (1564–1642) Newton (1643–1727) Figure 16.38 Time line of some famous early astronomers Although Kepler had found an accurate way to describe the motion of the planets mathematically, an explanation was not produced until about 80 years later (Figure 16.38) when Newton was able to use the motion of the planets and the Moon as evidence for his newly developed theory of universal gravitation (Chapter 6). 1 Many people would place Newton and Galileo in a list of the top five scientists of all time but, to a certain extent, that is just a matter of opinion. a Why do you think Newton and Galileo are so highly respected? b What criteria would you consider when trying to decide if a scientist was ‘great’? 2 Research the origins of the quotation ‘the shoulders of giants’, which forms the title of this Additional Perspectives section. The Big Bang model The Big Bang model is the current theory about how the universe began at one precise moment in time, 13.8 billion years ago. Before looking more closely at this theory, we will first consider the evidence for an expanding universe. Using spectra to determine the velocity of stars and galaxies If a source of light is not stationary but moving towards or away from an observer, there will be a shift (a very slight change) in all the wavelengths and frequencies of the light received. This is similar to the Doppler effect in the sound received from moving vehicles – as a police car approaches, we hear a higher-pitched sound (shorter wavelength) than when it is moving away from us (Figure 16.39). moving police car Figure 16.39 The Doppler effect for sound 16.3 (D3: Core) Cosmology  29 In the case of light waves, the shift is very small and is usually undetectable unless a source is moving very quickly, such as a star or galaxy. In order to detect a shift for light we need to examine the line spectrum from the source and compare it to the line spectrum produced by the same element(s) on Earth. We find that the pattern of the absorption lines on a spectrum is the same, but all the lines are very slightly shifted from the positions they would occupy if there were no motion of the source relative to the observer. Careful observation of the line spectrum received from a star (Figure 16.40) can be used to calculate the velocity of the star. In example A in Figure 16.40, all the absorption lines have been shifted towards lower frequencies and this is commonly described as a red-shift. A red-shift occurs in the radiation received from a star or galaxy that is moving away (receding) from the Earth. If a star or galaxy is moving towards Earth, then the shift will be towards higher frequencies and is called a blue-shift, as shown in example B. (This is unusual for galaxies.) lower frequency (red end of spectrum) ‘normal’ spectrum from a source which is not moving compared with the observer A red-shift from a star (or galaxy) moving away from us B blue-shift from a star (or galaxy) moving towards us Figure 16.40 Red- and blue-shifts For a given wavelength, λ0, in a line spectrum, the shift (difference) in wavelength, Δλ, received from a fast-moving star or galaxy is proportional to its speed towards or away from the observer. The ratio of Δλ/λ0 is the numerical representation of red-shift and is given the symbol z. For a speed v, which is significantly slower than the speed of light, c, the red-shift, z, is given by the equation z= Δλ v ≈ λ0 c V. I. Relation This equation is given in the Physics data booklet and is similar to the equation used in Chapter 9. Because it is a ratio, red-shift does not have a unit. If we can measure the red shift for a known wavelength, we can calculate the recession speed of the source (star or galaxy). Basic calculations like these assume that the source of light is moving in a straight line directly away from the Earth. As we shall see, this is a reasonable assumption, although it is not necessarily perfectly true. Worked example 7 A line in the hydrogen spectrum has a wavelength of 4.34 × 10 –7 m. When detected on Earth from a distant galaxy, the same line has a wavelength of 4.76 × 10 –7 m. What is the speed of the galaxy? Δλ = (4.76 × 10 –7) – (4.34 × 10 –7) = 4.2 × 10 –8 m Δλ v = z= c λ 0 v 4.2 × 10 –8 = 4.34 × 10 –7 3.00 × 108 v = 2.90 × 107 m s–1 Good example..... 16 Astrophysics Because the shift is to a longer wavelength (a red-shift), we know that the motion of the galaxy is away from Earth. We say that the galaxy is receding from Earth. When the light from a large number of galaxies is studied, we find that nearly all the galaxies are receding from Earth and each other. This can only mean that the universe is expanding. 47 What is the red-shift of a galaxy with a recession speed of: a 2.2 × 106 m s–1 b 10% of the speed of light? 48 What is the recession speed of a galaxy (km h –1) if radiation of original wavelength 6.5 × 10 –7 m undergoes a red-shift of 3.7 × 10 –8 m? 49 A star receding at a velocity of 9.2 × 103 km s–1 emits radiation of wavelength 410 nm. What is the extent of the red-shift of this radiation when it is received on Earth and what is its received wavelength? 50 Hydrogen emits radiation of frequency 6.17 × 1016 Hz. What frequency will be detected on Earth from a galaxy moving away at 1.47 × 107 m s–1? 51 Only a very tiny percentage of galaxies are moving towards us. Research the blue-shift of the Andromeda galaxy, one of the galaxies in the Local group. Hubble’s law In the mid-1920s, the American astronomer Edwin Hubble compared information about the recession speeds of relatively nearby galaxies (obtained from the red-shift of the light received) with the distances of the galaxies from Earth that were determined by using Cepheid variables within the galaxies. By 1929 Hubble had gathered enough data to publish a famous graph of his results for Cepheids within distances of a few Mpc from Earth. Figure 16.41 includes more results and for greater distances. Figure 16.41 Variation of recession speeds of galaxies with their distances from Earth Velocity of recession/km s–1 30 10 000 8000 6000 4000 2000 0 1 20 40 60 80 100 120 140 Distance/Mpc Even today there are significant uncertainties in the data represented on this graph (although error bars are not shown on Figure 16.41). These uncertainties are mainly because the precise measurement of distances to galaxies is difficult, but also because galaxies move within their clusters. Nevertheless, the general trend is very clear and was first expressed in Hubble’s law: The current velocity of recession, v, of a galaxy is proportional to its distance, d (from Earth). This can be written as: v = H0 d This equation is listed in the Physics data booklet. 16.3 (D3: Core) Cosmology  31 H0 is the gradient of the graph and is known as the Hubble constant. Because of the uncertainties in the points on the graph, the Hubble constant is not known accurately, despite repeated measurements. The currently accepted value is about 70 km s–1 Mpc–1 (this unit is more widely used than the SI unit, s–1). However, different determinations of the Hubble constant have shown surprising variations. Hubble’s ‘constant’ is believed to be a constant for everywhere in the universe at this time, but over the course of billions of years its value has changed. Worked example 8 Estimate the gradient of the graph in Figure 16.41 and compare it with the value given for the Hubble constant in the previous paragraph. v 9000 =    120 d = 75 km s−1 Mpc −1 gradient, H0 = This value varies by 7% from the value quoted earlier, but neither figure includes any assessment of uncertainty, so it is possible that they are consistent with each other. Hubble’s law can be applied to the radiation received from all galaxies that are moving free of significant ‘local’ gravitational forces from other galaxies. That is, the law can be used for isolated galaxies or clusters (considered as one object), but is less accurate for individual galaxies moving within a cluster because the resultant velocity of an individual galaxy is the combination of its velocity with respect to the cluster and the recession velocity of the cluster as a whole. A few galaxies even have a resultant velocity towards the Earth at this time and radiation received from such galaxies is blue-shifted. The use of the Hubble constant with the recessional speeds of distant galaxies provides astronomers with another way of calculating the distance to galaxies which are too far away to use alternative methods. More about the Big Bang The conclusion from Hubble’s observations can only be that the universe is expanding because (almost) all galaxies are moving away from the Earth. It is important to realize that this is true for galaxies observed in all directions and would also be true for any observer viewing galaxies from any other location in the universe. Almost all galaxies are moving away from all other galaxies. Our position on Earth is not unique, or special, and we are not at the ‘centre’ of the universe – the universe does not have a centre. Calculations confirm that the further away a galaxy is, the faster it is receding. This simple conclusion has very important implications: the more distant galaxies are further away because they travelled faster from a common origin. Observations suggest that all the material that now forms stars and galaxies originated at the same place at the same time. An expanding model of the universe had been proposed a few years earlier by Georges Lemaître and this was developed in the 1940s into what is now called the Big Bang model. If radiation from a star or galaxy is observed to have a blue-shift, it is because it is moving towards Earth. This is not evidence against the Big Bang model because such an object is moving within a gravitationally bound system (a galaxy, a cluster of galaxies or a binary star system) and at the time of observation it was moving towards the Earth faster than the system as a whole was moving away. For example, our neighbouring galaxy, Andromeda, exhibits a small blue-shift – it is moving towards us as part of its motion within our local group of galaxies, which is a gravitationally bound system. In the Big Bang model, the universe was created at a point about 13.8 billion (1.38 × 1010) years ago. At that time it was incredibly dense and hot, and ever since it has been expanding and cooling down. The expansion of the universe is the expansion of space itself and it should not be imagined as being similar to an explosion, with fragments flying into an existing space (void), like a bomb exploding. 32 16 Astrophysics It may be helpful to visualize the expansion of space using marks on a very large rubber sheet to represent galaxies. (Imagine that the sheet is so large that the edges cannot be seen.) If the sheet is stretched equally in all directions, all the marks move apart from each other. Of course, a model like this is limited to only two dimensions (Figure 16.42). The red-shift of light should be seen as a consequence of the expansion of space rather than being due to the movement of galaxies through a fixed space. Figure 16.42 An expanding universe before now later It is very tempting to ask ‘what happened before the Big Bang?’ In one sense, this question may have no answer because the human concept of time is all about change – and before the Big Bang there was nothing to change. The Big Bang should be considered as the creation of everything in our universe – matter, space and time. Nature of Science Simplicity Expressed in basic terms, the Big Bang model of the universe is elegant in its simplicity. In judging scientific theories and models, as well as other human endeavours, simplicity is often (but not always) an admirable aim. This has been expressed in what is known as Occam’s razor – if you need to choose between two or more possible theories, select the one with the fewest assumptions. Until you know that a more complicated theory is preferable, simplicity may be the best criterion to judge between opposing models. More complicated theories are more difficult to test and, if they are found to be in doubt, it is often possible to add another layer of (unproven) theory to retain some of their credibility. Age of the universe We can make an estimate of the time since the Big Bang (the age of the universe) using Hubble’s constant (70 km s–1 Mpc–1). Because time, t = distance, d/velocity, v and v = H0 d, we can write: T≈ 1 H0 where T is the approximate age of the universe, often called Hubble time. This equation is given in the Physics data booklet. Calculated using this equation, T can be considered to be an approximate and upper limit to the age of the universe for the following reasons. It is not sensible to assume that the recession speed of the galaxies has always been the same. It is reasonable to assume that the speed of galaxies was fastest in the past when they were closer together and that they are now slowing down because of gravitational attraction. (We now know that this is not true: discussed in more detail later.) We do not know that the expansion started at the same time as the Big Bang. The uncertainty in the Hubble constant is significant. 16.3 (D3: Core) Cosmology  33 We would prefer the time to be in SI units, and in SI units Hubble’s constant becomes: 70 × 103 = 2.27 × 10 –18 s 3.26 × 106 × 9.46 × 1015 so that 1 1 T= = = 4.4 × 1017 s (or 1.4 × 1010 years). H0 2.27 × 10−18 52 What is the recession speed (km s–1) of a galaxy that is 75 Mpc from Earth? 53 How far away is a galaxy travelling at 1% of the speed of light? 54 Galaxy A is a distance of 76 Mpc from Earth and is receding at a velocity of 5500 km s–1. Another galaxy, B, is receding at 7300 km s–1. Without using a value for H 0, estimate the distance to galaxy B. 55 A spectral line of normal wavelength 3.9 × 10 −7 m is shifted to 4.4 × 10 −7 m when it is received from a certain distant galaxy. a How fast is the galaxy receding? b How far away is it? Cosmic microwave background (CMB) radiation Figure 16.44 Spectral distribution for a temperature of 2.76 K Intensity Figure 16.43 ‘Ripples in Space’; a map of the whole sky at microwave wavelengths showing the very small variations (1 part in 100  0 00) in the CMB radiation – firstly from COBE satellite and, later, in more detail from the WMAP satellite When it was first proposed seriously in the late 1940s, many astronomers were not convinced by the Big Bang model. (They mostly preferred what was then known as the Steady State Theory of an unchanging universe.) However, the discovery in 1964 by Penzias and Wilson of cosmic microwave background (CMB) radiation provided the evidence that confirmed the Big Bang model for most astronomers. Penzias and Wilson discovered that low-level microwave radiation can be detected coming (almost) equally from all directions (it is isotropic), rather than from a specific source. (Later, important tiny variations were discovered in the CMB, a discovery that has vital implications for understanding the non-uniform structure of the universe and the formation of galaxies (Figure 16.43.) Cosmic background radiation has been a major area of astronomical research for many years, including by the Cosmic Background Explorer (COBE) satellite. A very large amount of data has been collected and analysed by astronomers from many different countries. We have seen before that everything emits electromagnetic radiation and that the range of wavelengths emitted depends on temperature. The Big Bang model predicts that the universe was incredibly hot at the beginning and has since been cooling down as it expands, so that the average temperature of the universe should now be about 2.76 K. Figure 16.44 shows the black-body radiation spectrum emitted from matter at 2.76 K. When this isotropic radiation was discovered by Penzias and Wilson coming (almost) equally from all directions, the Hot T = 2.76 K Big Bang model was confirmed. Wien’s law (for black bodies) can be used to confirm the peak wavelength associated with this temperature: 1.1 mm Wavelength 34 16 Astrophysics λmaxT = 2.9 × 10 –3 m K λmax = 2.9 × 10 –3 = 1.1 × 10 –3 m 2.76 An alternative (and equivalent) interpretation of CMB radiation is that the shorter wavelengths emitted when the universe was hotter have stretched out because of the expansion of space. The observable universe After the development of the Big Bang model it seemed that the universe could be finite, with a finite number of stars, each having a finite lifetime, thus limiting the amount of radiation that could reach Earth. More importantly, even if the universe is infinite, it was now known to have a definite age, which means that the universe that is observable to us is limited by the distance that light can travel in the time since the Big Bang. The universe that we can (in theory) observe from Earth is a sphere around us of radius 4.6 × 1010 ly. This is known as the observable universe or the visible universe. (This distance is longer than 1.4 × 1010 ly because space has expanded since the Big Bang.) If there is anything further away, we cannot detect it because the radiation has not had enough time to reach us. 56 Summarize the two major discoveries that support the Big Bang model of the universe. 57 Astronomers look for ‘shifts’ in spectra as evidence for an expanding universe. The spectrum of which element is most commonly used, and why? 58 Draw a diagram to help explain why the light from some galaxies may be blue-shifted. 59 How will the average temperature of the universe change in the future if: a the universe continues to expand b the universe begins to contract? The accelerating universe and red-shift (z) What happens to the universe in the future is obviously dependent on the rate at which it is expanding and whether or not the expansion will continue indefinitely. Previously it was believed that the receding galaxies were simply losing kinetic energy and gaining gravitational potential energy, like objects projected away from the Earth, and that the fate of the universe depended on their initial speeds and the mass in the universe. But in recent years it has been discovered that the rate of expansion of the universe is not decreasing, but increasing. This is discussed in more detail in Section 16.5 (Additional Higher). The evidence for the accelerating expansion of the universe comes from the observation of supernovae. When a certain kind (Type 1a) of supernova occurs, the energy released is always about the same and it is well-known to astronomers. This information can be used to determine the distance to such events using b = L/4πd2. This means that such supernovae can be used as ‘standard candles’ for determining the distances to distant galaxies. Work on this topic by three physicists, Perlmutter, Riess and Schmidt (Figure 16.45), was jointly awarded the Nobel prize for physics in 2011. The red-shifts from Type 1a supernovae have been found to be bigger than previously expected for stars at that distance away, strongly suggesting an ‘accelerating universe’. This, of course, requires a new explanation and astronomers have proposed the existence of dark energy, a form of energy of low density, but present throughout the universe. Again, this will be discussed in Figure 16.45 Adam Riess, Saul Perlmutter and Brian more detail in Section 16.5. Schmidt 16.3 (D3: Core) Cosmology  35 Cosmic scale factor, R Astronomers use the cosmic scale factor to represent the size of the universe by comparing the distance between any two specified places (two galaxies, for example) at different times. These distances, and the cosmic scale factor, increase with time because the universe is expanding. cosmic scale factor (at a time t), R = separation of two galaxies at time t separation of the same two galaxies now Figure 16.46 Possible futures for the universe Cosmic scale factor, R (relative size of the universe) Because it is a ratio, the cosmic scale factor does not have a unit. It varies with time. From the definition, it should be clear that at this time R = 1, in the past R < 1 and (in an expanding universe) in the future R > 1. If at some time in the future the universe has doubled in size, R will equal 2 at that time. More generally, we can define the cosmic scale factor as follows: separation of two galaxies at time t cosmic scale factor (at a time, t), R = separation, d0, of the same two galaxies at a specified time, t0 d(t) R(t) = d0 Figure 16.46 shows some predictions for the possible size of the universe in the future (and how it might have been in the past). The red line represents an accelerating universe. This will be discussed in more detail in Section 16.5. The blue line represents a universe that will continue to expand for ever (but at a decreasing rate). The green line represents a universe that will expand for ever but at a rate that reduces to zero after infinite time. The orange line represents a universe that will reach a maximum size and then contract. 4 3 2 1 0 –10 Now 10 20 30 Time/billions of years Relationship between red-shift and cosmic scale factor We know that: Δλ λ – λ0 = λ0 λ0 where λ is the wavelength received from a distant galaxy because of the expansion of space, and λ0 is the wavelength that was emitted. Because the expansion of the wavelength can be represented by an increase in the cosmic scale factor between the time the light was emitted, R0, and the time it was received, R, we can write: λ – λ0 R – R0 = z= λ0 R0 red-shift, z = 36 16 Astrophysics or: z= R −1 R0 This equation is given in the Physics data booklet. Worked example 9 The light from a distant galaxy was found to have a red-shift of 0.16. a What was the recession speed of the galaxy? b Determine the cosmic scale factor when the light was emitted. c Estimate the size of the observable universe at that time (size now = 4.6 × 1010 ly). a v c 0.16 = z≈ v 3.0 × 108 v = 4.8 × 107 m s–1 R –1 R0 1 0.16 = –1 R0 V. V. Important b z= R0 = 0.86 c 0.86 × 4.6 × 1010 = 4 × 1010 ly 60 a Explain what is meant by the term ‘standard candle’. b Why are observations of supernovae considered to be the best way of determining the distances to remote galaxies? 61 Suggest what future for the universe is represented by the orange line in Figure 16.46. 62 Measurements of the light from a distant galaxy show that a line on its spectrum is 4.8 × 10 –8 m longer than when measured on Earth. If the light was emitted with a wavelength of 6.6 × 10 –7 m, a what is the value of the red-shift? b Calculate the cosmic scale factor at the time the light was emitted. ToK Link The history of astronomy has many paradigm shifts A paradigm is a set of beliefs, or patterns of thought, with which individuals or societies organize their thinking about a particular issue, whether it is big or small. It is like a framework for all our thoughts and actions when, for example, we try to understand how electricity flows down a wire, or decide which foods are healthy to eat. In scientific terms, a paradigm could be said to be a pattern of beliefs and practices that effectively define a particular branch of science at any period of time. An obvious example from this chapter would be the set of ideas associated with the, now discredited, belief that the Earth is at the centre of the universe and the various consequences of that fundamental idea. The phrase paradigm shift has been used increasingly during the last 50 years since it was first popularized by Thomas Kuhn and others in the early 1960s. It is used especially with respect to developments in science. There are plenty of examples which suggest that, while scientific understanding, knowledge and practices obviously evolve and, hopefully, improve over time, many of science and technology’s greatest achievements have occurred following a relatively sudden (and perhaps unexpected or even seemingly unimportant) discovery or invention, or following the genius of an individual who has the insight to look at something in a completely new way. The phrase ‘to think outside the box’ has become very popular in recent years and it neatly summarizes an encouragement to look at a problem differently from the way others think about it (the ‘box’ being the paradigm). A paradigm shift occurs when new insights, technology and discoveries have such a fundamental effect that current ideas or beliefs have to be rejected. Most individuals, organizations and societies find that a 16.4 (D4: Additional Higher) Stellar processes  37 very difficult

Use Quizgecko on...
Browser
Browser