SPECIAL SPECIALSECTION: SECTION:SOLAR SOLAR PHYSICS The physics laboratory in the sky E. N. Parker Enrico Fermi Institute, University of Chicago, 1323 Evergreen Road, Homewood, Illinois 60430, USA The Sun, by virtue of its large mass, size, and temperature, exhibits a variety of effects that are unknown in the terrestrial laboratory, thereby challenging the physicist to relate them to the basic principles of physics derived from the terrestrial laboratory. A number of solar phenomena are reviewed, with comments on the degree to which they are presently understood. 1. Introduction THE Sun has played a prominent role in the development of physics since the times of Kepler and Newton, providing a window into phenomena unknown to the restricted scale of the terrestrial laboratory1. The Sun continues in that role today, providing mystery and opportunity in such diverse fields as lepton physics and magnetohydrodynamics. It challenges the experimental physicist to develop increasingly sophisticated tools for investigation and the theoretical physicist to see both new facets and extensions to basic laws of physics. With its large size and mass, it serves in two distinct roles: (i) as a passive self-gravitating thermonuclear object providing opportunities for testing existing laws of physics on a grand scale and (ii) as a large radiative, magnetic, fluid dynamics laboratory where nature displays astonishing dynamical phenomena, previously unknown and wholly outside conventional wisdom. The precision measurements of gravitational time delays in radio signals propagating close to the Sun are an example of the former, the sunspot is an example of the latter. To begin with a review of past triumphs, recall that the gravitational field of the Sun is responsible for the motions of the planets in the manner described by Kepler’s laws. The theory of mechanics and gravitation was put forth by Newton in 1686 with confidence because the theory described the motions of the planets, the Moon, etc. with remarkable precision. Over the following two centuries, observations of the motions of the planets were refined, using the Newtonian theory of mechanics and gravitation to compute the slight gravitational effects of the planets and their moons on each other, and including the precession of the elliptical orbits, the precession of the planetary spin axes, etc. By the middle of the 19th century, the observations had become so precise that it was possible to show that the precession of e-mail: parker@odysseus.uchicago.edu CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 the orbit of Mercury is of the order of 30 arc seconds per century greater than predicted by the Newtonian theory. By the end of the 19th century, the extra precession was established as close to 43 arc seconds per century. The only planet showing a significant anomaly was Mercury, closest to the Sun, where the presumably radial gravitational field of the Sun is strongest. Simon Newcomb suggested that the inverse square law of gravity might have to be modified. The modification of Newtonian theory in 1916 by Albert Einstein’s geometrized reformulation of gravitational theory, known as general relativity, brought theoretical mechanics into agreement with the observation by providing a theoretical precession larger than the Newtonian value by precisely 43 arc seconds per century. General relativity also predicted the small deflection of starlight passing near the Sun to be 1.75 arc seconds for a light ray grazing the surface of the Sun and quite different from the prediction of Newtonian gravity and special relativity theory. It also predicted a longer transit time for light or radio signals propagating across the deep gravitational potential well close to the Sun. Both of those effects have since been verified, the latter with great precision. Now an obvious question concerns the uniqueness of Einstein’s geometrical formulation of gravitational theory, based on the simplest lowest order terms of the metric tensor. There is a variety of other mathematical formulations that, like general relativity, reduce exactly to Newtonian gravitation in the limit of weak fields. They are in some ways more complicated than general relativity, involving an admixture of scalar field along with the tensor field. The Dicke–Brans–Jordan theory represents what was perhaps the most popular alternative. The theory introduces an extra free parameter defining the relative strength of the scalar field, thereby providing a wide range of precession of the perihelion of the orbit of Mercury. This is a qualitative difference from the unique value of 43 arc seconds per century predicted by general relativity. Dicke then pointed out that there is a possibility that the inner core of the Sun is spinning much more rapidly than the visible surface. The idea is based on the fact that newly-formed stars are observed to rotate rapidly, with periods of just a few days, as distinct from the 25-day period for the equatorial surface of the present middle-aged Sun (4.6 × 109 years). The initial angular momentum of a star is then reduced over the first 3 × 108 years of life by the mass loss (stellar wind) flowing out through the long arms of the extended magnetic field of the star. There is no reason to think that the magnetic field 1445 SPECIAL SECTION: SOLAR PHYSICS penetrates through the core of the star, so the process might well leave the core with the original spin. The point of this speculation is that a rapidly spinning core would be slightly flattened (oblate) by the centrifugal force, and the gravitational field of the core would be correspondingly nonspherical. It is readily shown that a consequence would be a precession of the perihelion of Mercury produced by the core alone. So, the error in the Newtonian theory would be less than the 43 arc seconds per century, and the unique prediction of general relativity would be in error, agreeing with the observational result only by chance coincidence. The experimental test of this idea lay in searching for the associated very slight oblateness of the surface of the Sun. Dicke developed an optical apparatus for scanning around the rim of the Sun to see whether it might be slightly oblate. The rim of the Sun is distorted by prominences, faculae, the chromosphere, etc. However, by scanning just inside the limb of the Sun, Dicke devised ingenious ways of reducing the contributions of these effects, and, in the end, there has been no conclusive demonstration of any interesting deviation from a purely circular solar disk. What is more, recent probing of the rotation of the Sun by helioseismology – on which more will be said later – shows no rapidly spinning core. The verdict seems to be that the Sun is round, general relativity is precisely correct, and there is no significant presence of a scalar gravitational field in addition to the tensor field of general relativity. Note, then, that general relativity has become the theoretical basis for understanding the dynamics of the expanding universe, with recent observational indications that the expansion may be accelerating and the bothersome cosmological constant may not be identically zero. 2. Internal structure Now if the gravitational field of the Sun and the rotation rate of the core have played a central role in establishing the correct formulation of gravity and mechanics, what can the structure and behaviour of the solar interior tell us about its physics? The development of the kinetic theory of gases and the recognition of the critical temperature of liquids and solids in the second half of the 19th century led to the realization of the entirely atomic gaseous composition of the Sun, too hot throughout the interior for any chemical bonding to form molecules. Spectroscopy and thermodynamics established the surface temperature at 5600 K. The virial theorem of mechanics made it clear that the interior must be very much hotter (∼ 107 K) in order to maintain the radius (7 × 105 km) in opposition to the immense gravity (28 times the gravitational acceleration of 9.8 m/sec2 at the surface of Earth). Lane in 1869 and Emden in 1907 constructed simple mathematical models of the interior of the Sun by assuming that the temperature varies as some power of the density, T ∼ ργ–1, where γ (> 1) is a constant. Thus, for instance, 1446 if a purely monatomic gas were to mix adiabatically throughout the interior, γ would be 5/3. The essential point is that the mathematical model could be made to fit the mass and radius of the Sun by suitable choice of mean atomic weight and γ; and one could see how the kinetic theory of gases and Newtonian gravitation provide a satisfactory representation of the Sun. To do more required some knowledge of the physics of the outward radiative transfer of heat within the Sun2. An obvious question, then, was the heat source that maintains the temperature of the Sun. Chemical combustion is entirely inadequate, providing the energy of the Sun for only a couple of thousand years. Helmholtz in 1854, and later Kelvin, pointed out that the gravitational energy of the Sun is more than a thousand times greater than any hypothetical chemical energy. The concept is quite simple. The surface of the Sun is hot, so that it cannot avoid radiating energy away into space. So the internal thermal energy would decline except for the fact that gravity would cause the Sun to shrink, compressing the gas until it is even hotter than before in order to oppose the increased gravity of the shrinking Sun. So the temperature within, and at the surface, increases as the radius of the Sun declines. To put it in other words, the Sun is stable against radial disturbances. An inward kick merely causes the Sun to contract and then rebound. So the net result of the continuing radiation from the surface of the Sun would be slow contraction over several million years, with the gravitational force compressing and heating the gas to even higher temperatures. The Sun would grow brighter with the passage of time as a consequence of the temperature increase. The rate of contraction of the Sun is easily computed to be about 50 m/year and is entirely undetectable in the brief span of human science. The Sun would gradually increase in brightness over a couple of million years. It is interesting to note that Kelvin took this scenario so seriously that he rejected Darwin’s proposal of biological evolution on the grounds that there simply was not enough time for evolution to take place. Fortunately the geologists stepped in with clear evidence that – Darwin or not – the rock structures found at the surface of Earth require 108 to 109 years for their formation, and there was evidently plenty of time for biological evolution. Thus the age of Earth vitiated Kelvin’s gravitational contraction as the principal source of energy for the Sun. With the recognition by Einstein of the equivalence of matter and energy, thinking turned to the annihilation of mass by some unknown process as the sustaining source for the Sun2. With the luminosity of 4 × 1033 ergs/sec, this means that the mass of the Sun is declining at the rate of 4 × 1012 g/sec (4 million tons/sec). Ultimately, Bethe and others3–5 supplied the answer based on the newly founded nuclear physics, pointing out both the proton– proton chain and the carbon cycle. Both processes fuse four hydrogen nuclei into one helium nucleus, converting CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS about one per cent of the mass into energy. The proton– proton chain is the more effective of the two in the Sun, where the central density is of the order of 102 g/cm3 and the temperature is 1.5 × 107 K. Hydrogen is converted into helium at the rate of 4 × 1014 g/sec (400 m tons/sec). Now the present age of the Sun is inferred from the assumption that the Sun and Earth were formed more or less simultaneously, and the uranium and lead isotope ratios in the crust of Earth indicate an age of 4.6 × 109 years. In that span of time, approximately 6 × 1031 g of hydrogen, representing 3 per cent of the mass of the Sun, have been converted to helium, with a total loss in solar mass of 6 × 1029 g. The production of helium in the central core of the Sun increases the mean atomic weight, of course, and the result is a slow contraction of the core to increase the temperature to support the core in opposition to gravity. The helium also has the effect of diluting the hydrogen so that the core contracts until the density and temperature increase to maintain the thermonuclear energy supply. Thus the Sun is now smaller, hotter, and approximately 30 per cent brighter than it was in early times. The fact that the Sun was substantially fainter 4 × 109 years ago raises interesting questions about the temperate climate of Earth, and even Mars, at that time. 3. Composition of the Sun Before going into helioseismology and neutrino emission, consider the knowledge of the interior of the Sun about a hundred years ago. Spectroscopy showed a spectrum dominated by the lines of such elements as carbon, calcium, sodium, silicon, iron, magnesium, etc., suggesting that the Sun was composed largely of the vapours of these substances. The immediate problem was that such large atomic weights required a much hotter Sun than the mathematical models could accommodate to maintain the Sun against its own gravity. Hydrogen and helium (which was discovered spectroscopically on the Sun before it was known in the laboratory), suggested by the atomic weights needed for the mathematical models, were relatively inconspicuous in the solar spectrum, and in any case were too transparent at 5600 K to account for the opaque surface layer of the Sun. The fuzziness of the visible surface of the Sun in white light is limited to something of the order of 100 km at the center of the disk. To make a long story short, the dilemma was resolved by the negative hydrogen ion—a hydrogen atom with an additional electron orbiting about it. Hans Bethe first calculated the existence of this atomic structure, pointing out in 1929 that the electron is attracted to the electrically neutral hydrogen atom because the presence of the external electron attracts the positively charged nucleus and repels the negative electron cloud. Thus, the otherwise spherical hydrogen atom is slightly distorted with a small net positive CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 charge in the end pointing toward the external electron. Ten years later, Wildt6 pointed out the astrophysical implications of the negative hydrogen ion. The loosely bound external electron is nicely tuned to absorbing and reradiating light, so it serves very well at the 5600 K temperature of the visible surface of the Sun in impeding the passage of light and providing a relatively sharp visible surface. Chandrasekhar7 and others undertook the difficult task of precise calculations of the properties of the negative hydrogen ion. Spectroscopic observation of the ‘red flames’ or chromosphere of the Sun that peek around the limb of the Moon during an eclipse indicated that hydrogen is the most abundant element in the Sun. Hydrogen shows clearly in the chromosphere because the temperatures are somewhat higher (6000–8000 K) than at the visible surface. The 5600 K at the surface only weakly excites the sturdy hydrogen atom, where the hydrogen is able to hang on to that extra electron to form the negative ion. The chromosphere is better suited to exciting the hydrogen atom, but, of course, is so hot as to knock the negative hydrogen ion to pieces. So the dilemma was resolved! The Sun is mostly hydrogen, with about one in ten atoms being helium. The heavier elements are present only at the one per cent level, or less, and their weaker electron structures are strongly excited by the 5600 K at the surface so that they dominate the line spectrum there. With some knowledge of the composition of the Sun, the next step in working out the structure of the interior requires knowledge of the impediment of the gas to the passage of electromagnetic radiation, dominated by soft X-rays in the core where T ∼ 1.5 × 107 K, and by visible light at the surface where T ∼ 5600 K. There is an ultraviolet domain between. The impediment is called the opacity and its calculation requires computing the absorption and reradiation by each of the atomic constituents. The high temperature in the deep interior of the Sun means that the atoms are highly ionized, with only the heavier elements, e.g. Ca, Si, Fe, etc. able to hang on to one or two of their innermost electrons. These few bound electrons make a major contribution to the opacity so essentially all atomic species have to be treated in their several states of ionization. One can begin to see the immensity of the task2: One needs precise detailed models of the solar atmosphere in order to determine the relative abundances of the elements at the visible surface from spectroscopic observations. The assumption is made that the relative abundances of the elements is the same in the interior as at the surface. Then the enormous calculation begins computing the probability of each number of electrons knocked off each different element at arbitrary temperature and density, and then computing the absorption of radiation by the electrons still attached. From this a table of opacity as a function of temperature and density can be constructed. Then finally, one can address the problem of computing inward from the 1447 SPECIAL SECTION: SOLAR PHYSICS visible surface where the temperature and density are known. The temperature must increase inward sufficiently rapidly to provide the observed outflow of heat (4 × 1033 ergs/sec). At the same time the density must increase inward sufficiently rapidly that the gas pressure (the product of temperature and density) is sufficient to support the weight of the overlying layers of gas against the gravitational field of all the matter below. For this reason, the procedure is usually reversed, starting with an assumed temperature (∼ 1.5 × 107 K) and density (∼ 100 g/cm3) at the center and calculating outward. The local rate of thermonuclear energy production has to be worked out too, of course, in order to know the outflow of energy at each radius. When the calculation arrives at the surface, where the temperature and density essentially vanish, the total mass and radius of the theoretical model are obtained. The calculation is repeated for different central temperature and density until the result matches the Sun. Then, once the Sun is properly modelled, the calculations can be applied to other stars where the radius is not precisely known, although the masses can be determined in double-star systems. It is interesting to note, then, that the deep interior of the Sun turns out to be stably stratified against vertical mixing. That is to say, the central core is hotter than the overlying layers, but not enough hotter that it can exchange places. Any upward displacement of gas from the core provides expansion and cooling to such a degree that the uplifted gas is cooler than the ambient gas, so that the uplifted gas is denser than its surroundings and relaxes back into the core. This is all based on well-known principles of physics, but the computation of opacities, with hundreds of different ions, is so complex and tedious that estimates are introduced in place of detailed computations for many classes of ions. The problem has been taken up in weapons laboratories where the properties of nuclear bomb explosions are explored theoretically, using many of the same opacities. The tables of opacities have been improved for this purpose, allowing more precise calculation of the interior of the Sun. A number of refinements have been introduced, such as the accumulation of helium throughout the central core as the hydrogen burning progresses, and the gravitational settling of the heavier ions relative to the lighter constituents with the passage of time. The result is a detailed quantitative model of the internal structure of the Sun. Then comes the ultimate question: Is there some way to check this sophisticated theoretical model of the interior to be sure that it is correct? Are the abundances of the elements really the same as at the surface? Is there really no vertical mixing of the elements in the core? Is the Sun really as old, or as young, as the Earth? 4. Probing the interior The first test for the theoretical model of the solar interior 1448 was based on the emission of neutrinos from the thermonuclear reactions in the core. Neutrinos pass so freely through matter that they easily escape from the core, being the only messengers coming directly from the core to the outside world. Needless to say, their free passage through matter makes them extremely difficult to detect – difficult but not impossible8,9. With six-hundred tons of cleaning fluid in a tank deep in the Homestake Mine in South Dakota (to get away from the spurious effects produced by cosmic rays) Ray Davis looked for neutrinos above 0.8 MeV reacting with the chlorine in the cleaning fluid to produce the radioactive isotope 40A. The argon atoms are radioactive with a half-life of about a month, and are swept out of the cleaning fluid about once a month to be detected and counted by their radioactive decay. To make a long story short, the theoretical models of the solar interior forty years ago when the planning of the experiment was initiated, suggested one or two argon atoms per day. In fact Davis found nothing beyond statistical fluctuations in the background count rate. A careful re-evaluation and refinement of the theoretical model, leading to a small downward readjustment of the central temperature of the Sun, drastically reduced the expected neutrino detection rate, to something of the order of five per month. The continuing accumulation of data in the Homestake detector gradually reduced the statistical uncertainties, to where it began to appear that there really were neutrinos from the Sun, but only about a third of the expected number10,11. The question was whether the theoretical model of the Sun was in error, or was there something about the physics of neutrinos that was not properly understood? In particular, neutrinos have conventionally been assumed to have no rest mass, in the same way that a photon of light has no rest mass. That is to say, a neutrino travels through free space at the speed of light and experiences no passage of time. However, it is not impossible that a neutrino has a rest mass, so that it experiences the passage of time. If so, it would open up the possibility that the electron neutrinos emitted from the core of the Sun oscillate through the mu-neutrino state and tau neutrino state during their passage to Earth. In that way the electron neutrinos would spend only one-third of their time in the electron–neutrino state to which the Homestake detector is sensitive. The result would be the detection of only one third of the expected number of neutrinos. This dilemma has been cleared up in favour of unknown neutrino physics, because of the precise probing of the interior of the Sun through helioseismology. It was first noted by Leighton that the surface of the Sun is continually agitated by fluctuations with periods of the general order of 5 min. Imagine a pan of water sitting in a metal sink with some such machinery as a garbage grinder running immediately below. The vibrations set up a pattern of small-scale waves on the surface of the water. Roger Ulrich12 and Leibacher and Stein13 recognized that the oscillations in the Sun represent sound-waves CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS trapped between the surface of the Sun and a refractive turn-around far below the surface. The Sun rings like a bell, but at a thousand different frequencies all at once. The essential point is that the period of each mode of oscillation is just the sound-transit time along the path of the sound wave down and back up through the interior. Each acoustic mode represents a sound wave travelling on a different path down into the Sun, around, and back up to the surface. Some paths go deep, some are only shallow, etc. The theoretical model of the solar interior provides the speed of sound as a function of depth from which one can calculate the path of each acoustic wave mode, and the transit time along that path for each acoustic mode. The transit time determines the period of the oscillation at the surface of the Sun. The calculated periods are then compared to the observed periods for several hundred different modes, sampling different depths. It is a precision test of the theoretical model, and it is gratifying to find that, when all the refinements of the theoretical model are included, the periods provided by the theoretical model agree closely with the observed periods, indicating that the speed of sound in the theoretical model nowhere differs by more than onepart in five-hundred from the actual speed of sound— approximately the expected observational error11. It appears then that the theoretical model of the interior of the Sun is properly constructed and there are no anomalous abundances of elements in deeper layers of the Sun. This is a major triumph for theoretical physics. Turning again to the low level of neutrino emission from the Sun, the evident accuracy of the theoretical model of the interior leaves only unknown neutrino physics as the explanation for the discrepancy11. This challenge has generated a vigorous response among physicists. The first step was to build a neutrino-detecting system using gallium instead of chlorine (in both Italy and the former Soviet Union) to detect neutrinos down to 0.2 MeV, thereby picking up the neutrinos emitted in the initial p–p reaction of the proton chain. The huge water–Cerenkov detector, Kamiokande, in Japan, was also applied to the task, providing both the energy and direction of the incoming neutrino. The data is from these three additional projects, and they all detect neutrinos from the Sun, but only at about a third or half of the predicted number, more or less along the same lines as the Homestake detector. The next generation of neutrino detectors is coming on line, designed to give the energy and direction of each incoming solar neutrino at substantial count rates of ten or more per day (Super Kamiokande in Japan; the heavy-water Sudbury Neutrino Observatory in Ontario (Canada), and Borexino (Italy)). In the meantime, Super Kamiokande has shown that neutrinos do, in fact, have a nonvanishing rest mass, based on measurements of the difference in the up-and-down fluxes of the neutrinos produced by cosmic rays colliding with the terrestrial atmosphere. The essential point is that the downward neutrinos are fresh from their creation in the CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 atmosphere overhead, while the upward moving neutrinos have passed through the solid Earth in a time of the order of 30 milliseconds, allowing time for conversion to neutrino states not detected by Super Kamiokande. 5. Convection and magnetic fields There is a branch of classical physics that is challenged to come up with new concepts, by the seemingly innocent fact that the theoretical model of the interior of the Sun provides an outer region of convective turnover, extending up to the visible surface of the Sun from a depth of 2 × 105 km (solar radius = 7 × 105 km)14. The continual turnover of the gas arises because the radiative transfer of heat deep in the Sun becomes less effective as the temperature diminishes outward, with the result that it cannot handle the heat transport when the temperature falls below about 2 × 106 K at the depth of 2 × 105 km. The outward heat flow is 4 × 1033 erg/s, and to transport so much heat by radiation the temperature would have to decline outward so rapidly that the cooler gas above becomes too dense, and the hotter gas below becomes too tenuous for stability. The hot and cold gases prefer to exchange places, providing convective turnover just like the rolling boil of the water in a pot on a hot stove. The resulting vigorous vertical mixing of the gas takes over the heat transport to the surface. The small effect of the convection on the overall theoretical model of the solar interior is easily handled by approximate methods known from hydrodynamics as the mixing length theory, and, as already noted, there seems to be no problem there. The new physics arises because the gas within the Sun is so hot as to be ionized, providing free electrons so that the gas is an excellent conductor of electricity – as good as cold copper in the central regions. The high electrical conductivity over the broad Sun means that there can be no significant electric field in the local frame of reference of the moving fluid, because any attempt to initiate an electric field would be met by a rush of free electrons, neutralizing the attempt. It follows that any magnetic field present in the gas is carried along bodily in the swirling convection, always moving in the frame of reference in which there is no electric field. The magnetic field is swirled, stretched, and deformed just like a wisp of smoke. Only when the magnetic field becomes so strong that it can physically stop the convection does the mixing, stretching, and intensifying of the field cease. The effects that arise from this simple magnetic transport property of hot ionized gas, or plasma, are legion, and involve hitherto unfamiliar combinations of hydrodynamics, magnetohydrodynamics, and local radiative transfer. First of all, there is the convection itself, extending from where the plasma density is 0.2 g/cm3 at the base of the convective zone up to the visible surface where the plasma density is 0.2 × 10–6 g/cm3. An understanding of convection in an atmosphere with such strong vertical stratification is only 1449 SPECIAL SECTION: SOLAR PHYSICS beginning to be developed. Numerical simulations are the principal tool for exploring the subject, but, computers are still some way from being able to handle so much stratification. The convection has a number of characteristics quite unlike convection in a pot of water, where the fluid has essentially uniform density. For instance, numerical simulations have shown that there are downward plunging clumps of cold dense gases. The nonuniform rotation of the Sun is presumed to be a direct result of the convection, but it has not yet been possible to simulate the convection with sufficient accuracy and detail to show how this works. As another example, the formation of a sunspot represents a systematic concentration of magnetic field driven by the convection in opposition to the enormous pressure of the magnetic field, and once again the convective mechanism is not understood. The explosive flare phenomenon is an example of magnetohydrodynamic interaction where at least the general principles seem to be in hand. Flares are observed on all scales from the largest (∼ 1032 ergs over dimensions of 104 km) down to the limit of detection (∼ 1025 ergs over 102– 103 km), rapidly converting magnetic free energy into hot plasma and fast particles. The basic effect appears to be rapid dissipation and reconnection of nonparallel magnetic fields. The discovery of this peculiar aspect of classical physics was motivated by the otherwise inexplicable explosive conversion of magnetic energy required to explain the flare15,16. That is to say, the physicist is lured onward by the mysteries presented by Nature, and the observations provide some hint as to the nature of the unknown physical effects. Rapid reconnection involves such diverse phenomena as plasma turbulence in the small and the Petschek effect in the large, and the subject is active today, with attention directed to the various forms the dynamics can take in three dimensions and to the acceleration of ions and electrons to very high energies in the central regions of the dissipation. The ubiquitous nature of rapid reconnection in the astronomical universe is demonstrated observationally by the widespread astronomical appearance of million degree tenuous plasmas and fast particles, and is to be understood in terms of the spontaneous appearance of discontinuities (intense current sheets) in any magnetic field embedded in a plasma undergoing slow continuous deformation17,18. The effect arises from the nature of the Maxwell stress tensor for the magnetic field. Each surface of discontinuity, or current sheet, becomes a site for resistive instabilities and rapid reconnection, i.e. explosive dissipation of magnetic free energy. It appears that this general theoretical property of the magnetic field is the major heat source responsible for the X-ray emitting corona of the Sun, on which more will be said later. This brings us to the fact that the outer atmosphere of the Sun – the corona, conspicuous when the dazzling disk of the Sun is obscured by the Moon during an eclipse – is heated to temperatures in excess of a million degrees. So 1450 high a temperature was suggested a hundred years ago by the great outward extension of the corona in opposition to the powerful gravitational field of the Sun (28 times the acceleration of gravity at the surface of Earth). Then in the 1930s the temperature was confirmed by some clever experiments in the terrestrial spectroscopy laboratory, showing that the observed spectral lines are from 10, 11, and more times-ionized iron, silicon, calcium, etc., occurring only at million degree temperatures19–21. The corona is so tenuous (typically 108 atoms/cm3 near the Sun) that it cools only relatively slowly by radiation, while it has an enormous thermal conductivity as a consequence of the high temperature and the associated 104 km/sec thermal velocities of the free electrons. Thus it is not surprising that the million degree temperature extends far out into space. The corona is strongly bound by gravity near the Sun, where the mean thermal energy is only about a tenth of the energy necessary to escape from the Sun. The outer regions, far from the Sun, are not strongly bound and escape into space. The surprise was that Newton’s equation of motion showed that there is no static equilibrium for such an atmosphere, the only steady state being gradual outward acceleration, from negligible velocity (∼ 1 km/sec) near the Sun to supersonic velocity (300– 1000 km/sec) at large distance. This is the origin of the solar wind, which drags the weaker magnetic fields of the Sun along with it, thereby filling all of interplanetary space with an outward sweeping spiral (because of the rotation of the Sun) magnetic field22. The outward sweeping wind and field push back the galactic cosmic rays to some degree, impact and agitate the magnetic fields of Earth, Jupiter, Saturn, etc., and generally determine the dynamical state of space throughout the solar system23. The outstanding question remaining is the heat source that creates the million degree temperatures around the Sun. 6. Magnetic fields of the Sun Ordinarily, in the terrestrial laboratory, we do not think of magnetic fields in association with gases; we think of iron magnets and coils of copper wire carrying electric currents (electromagnets). But in fact, as already noted, a gas sufficiently hot as to be fully ionized and an excellent conductor of electricity over the large dimensions appropriate to the Sun carries with it whatever magnetic fields happen to be present, on a more or less permanent basis. Rapid reconnection at incipient discontinuities is the one scheme that can quickly dissipate the magnetic field, by the simple expedient of creating very small scales in the structure of the field. But this rapid dissipation is effective only in the tenuous outer atmosphere of the Sun, and it is not obvious that it occurs in the tumbling convection below the surface of the Sun, where the gas is so dense as to respond only sluggishly to the magnetic forces. Magnetic fields were first established on the Sun by CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS George Ellery Hale24 when he observed the Zeeman splitting of some of the spectral lines in sunspots. The splitting indicated magnetic fields of 2000–3000 G in the dark central umbra of the spot. That is a very strong field, comparable to what can be produced in a strong electromagnet. Hale went on to point out that the magnetic field in the leading and following spots of each bipolar sunspot group had opposite sign, suggesting that the field emerges from one, arches over and returns back into the Sun in the other. He noted, too, that magnetic fields have opposite directions in the northern and southern hemispheres, and the whole magnetic system reverses with each successive 11-year sunspot cycle, which is now often referred to as the magnetic cycle since it is driven by the generation of magnetic field deep in the convective zone. Following World War II, the advent of electronics made it possible to develop a much more sensitive magnetograph, which soon detected magnetic fields of other stars and showed that there is a general background dipole magnetic field in the Sun of about 10 G, extending in at the north pole and out at the south pole, or vice versa in the next 11-year period25. The general field reverses near the maximum of the sunspot cycle when most of the sunspots are popping up within about 15° of the solar equator. The Babcocks also found that the active regions, in which the sunspots and large flares occur, lie in the midst of extended bipolar regions of ∼ 100 G. The bipolar character of these regions of magnetic activity indicates that the magnetic fields are part of a general intense east–west magnetic field somewhere deep in the convective zone. The individual bipolar active region is created by the upward bulging of a segment of that east–west field, the bulge having the form Ω of the capital omega. The first question is obviously the origin of the magnetic fields, for which the answer seems to be that the nonuniform rotation of the Sun (for instance, the equatorial surface of the Sun has a 25-day rotation period, while the polar regions rotate in approximately 30 days) continually shears and stretches out the dipole component of the magnetic field into an east–west field, while the cyclonic rotation of the tumbling convection raises and rotates Ω loops in the east–west field to reinforce the dipole field26–28. Why, then, are sunspots formed in the otherwise 100 G surface regions of bipolar field? There one can only say that ‘we do not know’. The magnetic field somehow interacts with the convection and the convection somehow interacts with the magnetic field to compress the field into the 2000–3000 G, that is observed. This all takes place in opposition to the enormous pressure of the 2000–3000 G field. It is observed that the tenuous gas trapped in the strong (100 G) bipolar fields of the active regions is heated to temperatures of 2–5 × 106 K, and along some bundles of field lines the density becomes so high (109– 1010 atoms/cm3) as to produce strong thermal emission of Xrays (107 ergs/cm2 sec). In contrast, the broad regions of CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 weak field (5–10 G) are heated to temperatures of 1.5 × 106 K and more, with densities limited to 108 atoms/cm3 by the free expansion of the gas to form the solar wind, as already noted. It appears at the present time that this expanding corona is heated largely by the microflares (1025–1028 ergs) among the small magnetic elements that appear at the surface of the Sun, while the dense X-ray corona is heated by even smaller flares (nanoflares, below the limit of detection by present instruments) arising in the spontaneous discontinuities created in the 100 G bipolar fields by the continual intermixing of the footpoints of the field at the convecting surface of the Sun18. Speaking of the small magnetic elements brings us to further subtleties in the mysterious behaviour of the Sun. About thirty years ago it became clear from quantitative observational studies that the magnetic fields of the Sun do not form a continuum at the surface of the Sun, but instead are made up of widely separated tiny concentra-ted magnetic flux bundles or magnetic fibrils of 1000–2000 G with diameters of the general order of 100 km. Each separate fibril is held in the grip of the surrounding gas, for otherwise it would expand and disperse. The typical observed mean field of 10 G is merely a measure of the spacing of the many unresolved individual magnetic fibrils. The 100 G regions have their fibrils more closely spaced. Only recently has it been possible to detect the individual fibrils and it is not possible to resolve their internal structure nor to study their interactions with ground-based telescopes because the atmosphere above the telescope limits the resolution to ∼ 300 km even under the best seeing conditions. One asks why the field is in this peculiar state with so much ‘unnecessary’ free energy. New fibrils continually bulge upward through the surface, forming small Ω loops and jostling against each other to provide the microflaring that seems to be the main energy source for the solar wind. These small-scale fibrils and bipoles appear over almost the entire surface of the Sun. These background fibrils show only modest variation with the 11-year cycle, suggesting a different origin from the main fibril fields of the active regions. A critical review of the standard explanation for generating the magnetic fields of the Sun, already mentioned, offers no enlightenment, and, in fact, turns up another serious puzzle. For the fact is that the generation of magnetic field by the cyclonic convection and nonuniform rotation of the Sun requires that the magnetic field diffuse across the surface of the Sun and the depth of the convective zone during the 11-year magnetic cycle. No adequate diffusion mechanism is known. We used to think that the turbulent convection mixes the magnetic field over these dimensions, the way turbulence mixes a puff of smoke throughout a room. But it is now clear that the mean magnetic field in the deep convective zone is at least 103 G, and much too strong to submit to the ‘indignity’ of turbulent mixing. It may be that the answer to 1451 SPECIAL SECTION: SOLAR PHYSICS the dilemma lies in the individual fibril being the basic magnetic entity, rather than the mean fields usually employed in calculating the generation of magnetic field. The fibrils are capable of rapid reconnection where they meet and may be transported more freely. But, all this has to be worked out quantitatively before any claim to understanding can be made. It is clear that the first step in studying the fibril nature of the magnetic field is to develop and construct a groundbased telescope that can resolve and study the detailed properties of the magnetic fibrils, as the principal players in the magnetic activity of the Sun. They appear to be the microscopic architects of sunspots, flares, coronal heating and the solar wind, and the generation of magnetic field itself, and yet we cannot see them clearly from Earth. The development of adaptive optics to correct for the blurring by the atmosphere above the telescope has now progressed to the point that the necessary resolution of 0.1″ (75 km), or better, should be possible. High-spectral resolution combined with the necessary rapid-observing cadence (∼ 10 sec) and the high-spatial resolution require a large aperture (∼ 4 m) to gather enough photons. It is an essential step if we are to advance the physics of the active Sun. In fact the implications of the variable magnetic activity of the Sun, the associated varying brightness of the Sun, and the resulting climatic effects here at Earth, together with the implications for the activity of all stars and the new physics to be learned, place the successful construction of such a solar telescope at the highest priority, which brings us to the last section of this brief review. 7. The terrestrial challenge NASA began monitoring the brightness of sunlight with absolute radiometers on orbiting spacecraft in 1978, with the startling discovery that the brightness varies by as much as 0.15 per cent with the 11-year variation of sunspots, flares, and general magnetic activity of the Sun. More recent monitoring of other solar-type stars shows that they do much the same, with one such star, ominously, showing a decline in brightness of 0.4 per cent over only six years as its activity tumbled to low levels29. Jack Eddy30 emphasized some years ago that the Sun was almost entirely without activity over the seventy-year period of 1645 to 1715, called the Maunder Minimum after its discoverer at the end of the 19th century. With modern 14C-production data, Eddy went on to show that the Sun went through another extended inactive period during the 15th century and a prolonged state of hyperactivity during the 12th century. Zhang et al.29 estimated that the brightness of sunlight was depressed by something of the order of 0.4 per cent during the Maunder Minimum and probably enhanced above normal by a comparable amount during the 12th century. Then Eddy noted that the mean annual temperature in the Northern Temperate Zone varied up and down 1–2°C, while 1452 tracking these variations in solar activity. The close tracking of climate with solar activity has been investigated in detail since that time and proves to be much closer than Eddy could have imagined with the data available at that time. Historically, the extreme cold periods had devastating consequences for agriculture in northern Europe and China, and the warm periods had devastating consequences around the periphery of desert regions, e.g. what is now southwestern United States. The bottom line is that the great physics laboratory in the sky not only extends our opportunities to study physics, but some of its more mysterious demonstrations have profound implications for the human population of Earth. In particular, the Sun has become substantially more active during the 20th century, and presumably brighter on the average by as much as 0.1 per cent. The general warming of the climate from 1900 to 1950 would appear to be a consequence of this phenomenon. Since that time the climate picture has been complicated by the substantial increase in carbon dioxide in the atmosphere and the warmer sea water temperatures which discourage the absorption of the carbon dioxide into the oceans. It is a problem that needs to be thoroughly investigated so that we can have some idea of how to respond to these changes. 1. Parker, E. N., Solar Phys., 1997, 176, 219. 2. Eddington, A. S., The Internal Constitution of the Stars, Dover, New York, 1926. 3. Bethe, H. A., Phys. Rev., 1939, 55, 434. 4. Bethe, H. A. and Critchfield, C. L., Phys. Rev., 1938, 54, 248. 5. Weizsacher, C. F., Phys. Z., 1938, 39, 633. 6. Wildt, R., Astrophys. J., 1939, 89, 295. 7. Chandrasekhar, S., Astrophys. J., 1944, 100, 176. 8. Davis, R., Phys. Rev. Lett., 1964, 12, 303. 9. Bahcall, J. N., Phys. Rev. Lett., 1964, 12, 300. 10. Bahcall, J. N., Calaprice, F., McDonald, A. B. and Totsuka, Y., Phys. Today, 1996, p. 30. 11. Bahcall, J. N., Pinsonneault, M. H., Basu, S. and ChristensenDalsgaard, J., Phys. Rev. Lett., 1997, 78, 171. 12. Ulrich, R., Astrophys. J., 1970, 162, 993. 13. Leibacher, J. W. and Stein, R. F., Astrophys. J. Lett., 1991, 7, L191. 14. Schwarzschild, M., Structure and Evolution of the Stars, Princeton University Press, Princeton, 1958. 15. Parker, E. N., J. Geophys. Res., 1957, 107, 830. 16. Sweet, P. A., Nuovo Cim. Suppl., 1958, 8, 188. 17. Parker, E. N., Astrophys. J., 1972, 174, 499. 18. Parker, E. N., Spontaneous Current Sheets in Magnetic Fields, Oxford University Press, New York, 1994. 19. Grotrian, W., Naturwissenschaften, 1939, 27, 214. 20. Lyot, B., Mon. Not. R. Astron. Soc., 1939, 99, 580. 21. Edlen, B., Z. Astrophys., 1942, 22,30. 22. Parker, E. N., Astrophys. J., 1958, 128, 664. 23. Parker, E. N., Interplanetary Dynamical Processes, Interscience Div, J. Wiley and Sons, New York, 1963. 24. Hale, G. E., Astrophys. J., 1908, 28, 100, 315. 25. Babcock, H. W. and Babcock, H. D., Astrophys. J., 1955, 121, 349. 26. Parker, E. N., Astrophys. J., 1955, 122, 293. 27. Parker, E. N., Proc. Natl. Acad. Sci., 1957, 43, 8. 28. Parker, E. N., Cosmical Magnetic Fields, Clarendon Press, Oxford, 1979, pp. 532–815. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS 29. Zhang, Q., Soon, W. H., Baliunas, S. L., Lockwood, G. W., Skiff, B. A. and Radick, R. R., Astrophys. J. Lett., 1994, 427, L111. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 30. Eddy, J. A., Science, 1976, 192, 1189. 1453 SPECIAL SECTION: SOLAR PHYSICS Seismic sun S. M. Chitre and H. M. Antia Tata Institute of Fundamental Research, Homi Bhabha Road, Mumbai 400 005, India Even though the interior of the Sun is not directly accessible to observations, it is nonetheless possible to infer the physical conditions inside the Sun using the theory of stellar structure and the accurately measured frequencies of solar oscillations. The helioseismic data has provided a powerful tool to probe the Sun and also to test physical theories describing its internal constitution. 1. Introduction THE Sun has been verily described as the Rosetta Stone of astronomy. This is very apt since our nearest cosmic laboratory is readily available for studying a variety of physical processes operating both inside and outside the object. Astrophysicists have an abiding hope that the study of the Sun can serve as a guide for theory of structure and evolution of stars in general, and pulsating stars in particular. Clearly, its internal layers are not directly accessible to observations. Nonetheless, it is possible to construct a reasonable picture of the interior with the help of structure equations governing its equilibrium, together with the boundary conditions provided by observations. The principal question concerning the structure of the Sun is about checking the correctness of the theoretically constructed solar models. Fortunately, it turns out that the Sun is transparent to neutrinos released in the nuclear reaction network operating in the energy-generating core and also to seismic wave motions generated through the solar body. These complementary probes enable us to see inside the Sun and to infer the physical conditions prevailing in the solar interior and relate them to larger issues in astronomy and physics. The internal layers, in fact, provide an ideal celestial laboratory for testing atomic and nuclear physics, and high-temperature plasma physics and neutrino physics. on chain. The energy is transported outwards by radiative processes except in the outer unstable zone, extending over approximately a third of the solar radius below the surface where the energy flux is carried largely by convection modelled in the framework of a local mixing length theory. There is supposed to be no mixing of material outside the convection zone, save the slow gravitational diffusion of helium and heavy elements beneath the convection zone into the radiative interior, and there is no wave transport of energy or material. The standard nuclear and neutrino physics is adopted for constructing theoretical models satisfying the observed constraints, namely, Mass (M¤) = (1.9889 ± 0.0002) × 1033 g, Radius (R¤) = (6.9599 ± 0.0007) × 1010 cm, Luminosity (L¤) = (3.846 ± 0.006) × 1033 erg s –1, Age (t¤) = (4.6 ± 0.1) × 109 yrs, and Chemical composition (Z/X) = 0.0245 ± 0.002. (1) Here X and Z respectively, refer to the fractional abundance by mass of hydrogen and elements heavier than helium. The manner in which the pressure, density and temperature vary throughout the solar interior can be determined by solving the equations of mechanical and thermal equilibrium applicable to the spherically symmetric Sun1, where these variables are all taken to be functions of the radial coordinate, r. The structure equations are integrated numerically, with appropriate boundary conditions, with the auxiliary physical input of the opacity, nuclear energy generation rate and equation of state to construct solar models. The conventional approach to the theory of solar structure is to adopt at zero-age, a homogeneous chemical composition and the mass, and then to evolve the Sun over the solar age to yield the present luminosity and radius by adjusting the initial helium abundance and the mixing-length parameter which determines the convective flux in the convection zone. 2. Standard solar model 3. Seismic waves The standard solar model (SSM) is constructed using a variety of simplifying assumptions. The Sun is assumed to be spherically symmetric maintaining mechanical and thermal equilibrium, with negligible effects of rotation, magnetic field, mass loss or accretion of material and tidal forces on its overall structure. The energy generation takes place in the central regions by thermonuclear reactions converting hydrogen into helium mainly by the proton– proton chain. The energy is transported outwards by It has been observed since the early 1960s that the solar surface undergoes a series of mechanical vibrations; these manifest as Doppler shifts oscillating with a period centered around 5 min2. The pulsations have now been identified as acoustic modes of oscillation of the entire Sun3–5. Just like any musical instrument, the Sun also oscillates in a number of characteristic modes whose frequencies are determined by the internal structure and dynamics. The solar surface is seen to oscillate simultaneously in millions of modes, with the amplitude of an individual mode of the order of a few *For correspondence. (e-mail: chitre@astro.tifr.res.in) 1454 CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 Yl m (θ , φ ). SPECIAL SECTION: SOLAR PHYSICS centimeters per second. Remarkably, the frequencies of many of these modes have been determined to an accuracy of better than 0.01%. In much the same manner as the geophysicists are able to study the internal layers of Earth from seismic disturbances, the helioseismic tool furnished by the rich spectrum of velocity fields observed at the solar surface can probe the Sun’s internal layers to an extraordinary degree of precision. The accurately measured frequencies of oscillations, in fact, provide very stringent constraints on the admissible solar models. In order to determine the frequencies of these oscillations to high accuracy, one needs continuous observations extending over very long periods. From most observatories on the surface of Earth it is not possible to observe the Sun continuously for more than 15 h due to the day–night cycle. Thus, to get longer coverage of the Sun various strategies have been tried, which include observations from the geographic south pole, from a network of sites located around the Earth and observations from a suitably located satellite. There are several ground-based networks observing the Sun more or less continuously with a variety of instruments. The most prominent among these is the Global Oscillations Network Group (GONG) which includes six stations located in contiguous longitudes around the world6. GONG has been observing the Sun more or less continuously since 1995 and frequencies of approximately half a million modes have been calculated for different periods of observations7. Apart from earth-based networks, many instruments located on satellites have been observing the solar oscillations. The most important among these is the Michelson Doppler Imager (MDI) instrument8 on board the Solar and Heliospheric Observatory (SOHO) satellite, which was launched on 2 December 1995. The higher spatial resolution provided from space has enabled MDI to study oscillations with small-length scales. Solar oscillations may be regarded as a superposition of many standing waves, whose frequencies are controlled by the physical properties of the solar interior. There are two distinct types of wave-modes that the Sun can support: high-frequency acoustic modes (p-modes) for which pressure gradient provides the main restoring force; and low-frequency gravity modes (g-modes) for which buoyancy is the dominant restoring force, and separating these two classes of modes are the fundamental modes (f-modes) which are essentially the surface gravity modes. The eigenmodes of oscillations can be characterized by three quantum numbers: the angular degree, l; azimuthal order, m; and radial order, n. The oscillation amplitudes are small, and so they can be analysed using a linear perturbation theory. Further, since the Sun is spherically symmetric to a good approximation, the eigenmodes of oscillations can be expressed in terms of the spherical harmonics, Thus, for example, the radial component of velocity can be expressed as v ( r , θ, φ, t ) = v nl ( r ) Yl m (θ, φ) ei ω nlmt . CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 (2) Here r is the radial distance from the center, θ the colatitude and φthe longitude and ωnlm is the frequency of oscillations. It is often convenient to express frequencies in terms of mHz and define ν = ω/2πas the cyclic frequency. In absence of rotation and magnetic field, the frequencies will be independent of the azimuthal order m; the rotation and other symmetry-breaking forces lift this degeneracy giving rise to splitting of the modes for a given value of n, l. The mean frequency of a given multiplet νnl is determined by the spherically symmetric structure of the Sun, while the frequency splittings are determined by the rotation rate, magnetic field and other asphericities in solar interior. Extensive observations from GONG and MDI instruments have provided the mean frequencies of modes with degree, l, from 0 to 4000 and frequencies from 1 to 10 mHz (refs 7, 9–11). These include the p- and f-modes, while the low-frequency g-modes have not yet been unambiguously detected. The p-modes are believed to be excited by turbulent convection in the subsurface layers12. The propagation characteristics of these seismic waves are affected by sound speed in the solar material, which increases inwards due to rising temperature with depth. Thus, a wave excited near the surface and propagating inwards is refracted away from the radial direction until at some depth it suffers a total internal reflection and bounces back to the surface. Near the solar surface the waves tend to get reflected because of sharply declining density. In this way, the acoustic waves are trapped within a cavity and the wave may travel around the Sun several times, establishing a standing wave pattern, in the process providing a global diagnostic of the solar interior. Acoustic waves propagate through the body of the Sun along ray paths, as shown in Figure 1. The penetration depth of a given wave depends on its horizontal wavelength, or the angle of inclination to the radial direction – shorter waves (oscillations with large l) are confined within relatively shallow cavities below the surface, while the longer waves (small l) propagate deeper penetrating practically to the central regions. The radial modes (l = 0) propagate radially and hence suffer no refraction, thus penetrating all the way to the center. As different modes are trapped in different regions of solar interior, they sample properties of the region where they are trapped. This improves the diagnostic potential of solar oscillations since by studying the properties of a large variety of modes it is possible to infer the conditions over a sizeable fraction of solar interior. The disturbances observed at the photosphere naturally encounter the ‘murky’ surface layers which influence the oscillation frequencies to a significant extent; these surface effects must be properly filtered out while analysing the seismic data. Clearly, the characteristics of p-modes are mainly determined by the sound speed inside the Sun, but other properties like density, rotation velocity, magnetic field also affect the waves to smaller extent. Consequently, the accurately deter1455 SPECIAL SECTION: SOLAR PHYSICS mined frequencies of solar oscillations provide a powerful tool to probe the structure and dynamics inside the Sun. 4. Probes of the solar interior The initial attempts to learn about the solar interior were concerned with the boundary conditions at the surface. The spectroscopic data was extensively collected for studying the solar atmosphere, and the theory of solar structure was widely used to surmise the physical conditions below the surface for obtaining the observed temperature and luminosity. Since the 1960s, there have been valiant attempts to measure the flux of neutrinos generated by the nuclear reactions operating in the solar core (cf., Bahcall, this issue). The neutrino flux is sensitive to the temperature and composition profiles in the central regions of the Sun. It was, therefore, expected that the steep temperature dependence of some of the nuclear reaction rates will determine Sun’s central temperature to better than a few per cent. The persistent discrepancy between the measured solar neutrino counting rates and the predictions of standard models raised doubts about the reliability of structure calculations, based on the assumption of standard physical properties for neutrinos. This had prompted solar physicists to look for some independent means to explore conditions inside the Sun and the techniques of geo-seismology were adopted by using the precisely measured eigenfrequencies of global oscillations to determine the sound speed and density variations through most of the solar body. The helioseismic database of oscillation frequencies may be analysed in two ways: (i) forward method, and (ii) inverse method. In the forward method, an equilibrium solar model, constructed using the structure equations, is perturbed to obtain the eigenfrequencies of solar osci- llations in a linearized theory, and these are compared with the accurately measured oscillation frequencies. The fit is, of course, seldom perfect; but the comparison suggested that the thickness of the convection zone is close to 200,000 km, deeper than what was previously estimated, and helium-abundance by mass in solar envelope was indicated to be 0.25. The direct method has had only a limited success, though, since it is not possible to produce a perfect fit for the seismic data by merely fitting a set of adjustable parameters characterizing the specific models. As a result, the values for various parameters may be nonunique. A number of inversion techniques13 have, therefore, been employed to extract more information about the solar interior. One of the major accomplishments of the inversion techniques has been the effective use of the observed solar oscillation frequencies for a reliable inference of the internal structure of the Sun14,15. Thus, the profile of the sound speed, Γ1 = (∂ ln P/∂ ln ρ)S is the c = Γ(where 1P/ρ adiabatic index), has now been established through the bulk of the solar interior to an accuracy of better than 0.1%, and the profiles of a density, adiabatic index and other thermodynamic quantities are known to somewhat lower accuracy. It also appeared from the variation of sound speed beneath the convection zone that the adopted opacities for solar modelling near the base of the convection zone, were low by about 15–20%. This was later confirmed by the use of Livermore opacity calculations16. In Figure 2 are shown the plots of the relative difference in sound speed, and density between the Sun, as inferred from helioseismic inversions and a standard solar model with gravitational settling of helium and heavy elements17. There is a reasonably close agreement except for a noticeable discrepancy near the base of the convection zone and a smaller discrepancy in the energy-generating core. The bump below 0.7 R¤ could be attributed to a sharp change in the gradient of helium-abundance profile arising from diffusion in the reference model. A moderate amount of turbulent mixing (induced by say, a rotationally induced b instability) immediately underneath the convection zone can alleviate this discrepant feature. The dip in the relative sound speed difference around 0.2 R¤ is not yet well understood; it could be due to inaccurate composition profile in the solar model, possibly due to use of incorrect nuclear reaction rates. The sound speed profile in ionization zones is affected by the variation in Γ1 and it is possible to use this to determine the helium-abundance in solar convection zone. The inverted sound speed profile can be employed to compute the quantity, W (r ) = Figure 1. Propagation of acoustic waves corresponding to different values of l. The distance between two successive points at which a ray intersects the surface is a measure of its horizontal wavelength. Modes with low l or large horizontal wavelength penetrate into deeper layers. 1456 r 2 dc 2 , GM¤ dr (3) which is shown 3. The peakmodel around Figure in 3. Figure The function W(r)small for a solar is shown by the Figure b. Relative difference in sound speed and density continuous line, while the dashed line represents the same for the Sun r = 0.98 2Ra, in this curve is due to the HeII ioniza¤ 17 profiles between the inverted Sun and sound a standard solar model . using speed profile. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS tion zone, which can be calibrated to measure the helium abundance. The sharp change in its gradient around r = 0.713 R¤ indicates the base of the convection zone and this curve can thus be used to measure the depth of the convection zone. The helium abundance in solar envelope is found to be18 0.249 ± 0.003. This value is less than what was used in the earlier standard solar models and the discrepancy was attributed to the fact that some of the helium would diffuse into the interior through gravitational settling. An examination of the inverted sound speed profile, below the convection zone, also reinforces this conclusion19. The incorporation of gravitational settling in radiative interior, indeed, results in a significant improvement in solar models. This also reduces the life time of main sequence stars and the estimated age of globular clusters is also diminished. Clearly, this will have wider implications for the age problem in the context of standard big bang model of cosmology. The dip in Γ1 inside the ionization zone is also determined by the equation of state and the inverted sound speed in this region provides a test for the equation of state20. It is found that standard equations of state, which were widely used in stellar evolution calculations, were not good enough to model the solar interior. More sophisticated equations of state, like the MHD (Mihalas, Hummer and Dappen)21, or OPAL22 equation of state, are found to produce good accordance with helioseismic data. Further, the OPAL equation of state is found to be in better agreement with solar data compared to the MHD equation of state18. Even these equations of state show slight discrepancy in the core and this discrepency has recently been attributed to the neglect of relativistic correction for electrons23. In solar models the second derivative of temperature and hence that of the sound speed is discontinuous at the base of the convection zone. This discontinuity in the function W(r), eq. (3) can be utilized to identify the position of the base of the convection zone24. The sound speed as well as the frequencies of p-modes are very sensitive to the depth of the convection zone and therefore seismic inversions enable a very accurate measurement of its thickness. Using recent data the depth of the convection zone is estimated to be25 (0.2865 ± 0.0005) R¤. Further, the position of the base of the convection zone is controlled by the opacity of solar material. We can then estimate the opacity at the base of the convection zone26 and it has been found that the current OPAL opacity tables27 are consistent with helioseismic data to within an estimated error of 3%. The convective eddies inside the convection zone are expected to penetrate beyond the theoretical local boundary, but there is no satisfactory theory to describe this overshoot. A significant overshoot can alter the stellarevolution calculations and so far the extent of penetration is treated as a parameter in stellar evolution calculations. Now with the availability of helioseismic data, it has become possible to estimate this extent of overshoot below the base CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 of the solar convection zone. The discontinuity in the derivatives of sound speed at the base of the convection zone introduces an oscillatory component28 in the frequencies as a function of radial order n. The amplitude of this signal depends on the magnitude of discontinuity, which in turn depends on the extent of overshoot below the solar convection zone. Thus by measuring the amplitude of this oscillatory signal we can determine the extent of overshoot below the convection zone29,30. The measured oscillatory signal is found to be consistent with no overshoot and on the basis of this result an upper limit of 1/20 of the local pressure scale height has been obtained31 for the overshoot distance. This is, of course, too small to affect the stellar evolution calculations significantly. The frequencies of f-modes, which are surface gravity modes, are largely independent of stratification in the solar interior and are essentially determined by the surface gravity. These frequencies, which have now been measured reliably by GONG and MDI data, provide an important diagnostic of the near-surface regions, as well as an accurate measurement of the solar radius32,33. Furthermore, the changes in solar radius with time by a few km have also been recorded. Another application of the accurately measured f-mode frequencies is their potential use as a diagnostic of solar oblateness and of near-surface magnetic fields, in addition to the possibility of investigating solar cycle variation in these quantities. The primary inversions which have provided information about the physical quantities like the sound speed, density Figure 4. Fractional helium abundance by mass in the Sun as obtained from inversions is shown by the continuous line. The dashed line represents the abundance profile for a solar model without diffusion, the dotted line shows that for a model incorporating diffusion of helium and heavy elements. 1457 SPECIAL SECTION: SOLAR PHYSICS and adiabatic index in the solar interior are based on the equations of mechanical equilibrium. The equations of thermal equilibrium are not used, because on time scales of several minutes, no significant energy exchange is expected to take place in moving elements. The frequencies of solar oscillations are, therefore, largely unaffected by the thermal processes in the interior. However, having obtained the sound speed and density profiles in solar interior through primary inversions, we can employ the equations of thermal equilibrium to determine the temperature and chemical composition profiles inside the Sun34–36, provided input physics like the opacity, equation of state and nuclear energy generation rates are known. In general, the computed luminosity resulting from these inferred profiles would not necessarily match the observed solar luminosity. The discrepancy between the computed and measured solar luminosity can, in fact, provide a test of input physics, and using these constraints it has been demonstrated that the nuclear reaction cross-section for the proton–proton reaction, needs to be increased slightly to (4.15 ± 0.25) × 10–25 MeV barns36. This cross-section has a controlling influence on the rate of nuclear energy generation and neutrino fluxes, but it has never been measured in the laboratory and all estimates are based on theoretical computations. More recently, this cross-section has been revised upwards37 to a value close to what was estimated helioseismically. The inferred helium-abundance profile agrees with that in the standard solar model, incorporating diffusion of helium and heavier elements, except in layers just below the solar convection zone. This is the region where the solar rotation rate has a sharp gradient in radial direction (ChristensenDalsgaard and Thompson, this issue). The inferred heliumabundance profile, for example, shown in Figure 4 is essentially flat in this region. This indicates the presence of some sort of mixing process, possibly by rotationally induced instability which has not been properly accounted. The mixing in this region can also explain the anamolous low-lithium abundance in solar envelope. The destruction of lithium by nuclear reactions can take place at temperatures exceeding 2.5 × 106 K, but at the base of the solar convection zone the temperature is still not high enough to burn lithium. Thus, if the mixing extends a little beyond the solar convection zone to a radial distance of 0.68 R¤, the temperature can reach high enough value to explain the low abundance of lithium. This is exactly the region where the inferred composition profile is flat, indicating the operation of a mixing process. With an allowance of up to 10% uncertainty in opacity values, the central temperature of the Sun is found to be (15.6 ± 0.4) × 106 K (ref. 38). The inferred temperature and composition profiles may be used to compute the neutrino fluxes in the seismic solar models and the predicted neutrino fluxes come close to what is obtained for the current standard solar models. This suggests that the known discrepancy between the observed and predicted neutrino 1458 fluxes is likely to be due to non-standard neutrino physics. Thus, helioseismology has turned the Sun into a laboratory to study properties of neutrinos. Apart from spherically symmetric structure of solar interior, it is also possible to determine helioseismically the rotation rate inside the Sun from the accurately measured rotational splittings (cf., Christensen-Dalsgaard and Thompson, this issue), Dnlm = (νnlm–νnl – m)/2m. It turns out that first-order effects of rotation yield splittings which depend on odd powers of m, and these odd splitting coefficients have been used to determine the rotation rate as a function of depth and latitude. On the other hand, magnetic field or other asphericities give only even order splittings and hence these can be separated from the rotational effects. These even splitting coefficients allow us to study the departures from spherical symmetry inside the Sun. Further, the local helioseismic techniques (cf., Kosovichev and Duvall, this issue) allow us to study other large-scale flows, including meridional flows in solar interior. The Sun’s oblateness has been measured to be about 10–5 at the solar surface39, and there does not seem to be any evidence of temporal variation in oblateness. The oblateness is indeed, consistent with what is expected from the helioseismically inferred rotation rate in solar interior. The resulting quadrupole moment40 turns out to be (2.18 ± 0.06) × 10–7, which yields a precession of perihelion of planet Mercury’s orbit by about 0.03 arc sec/century, validating the general theory of relativity. The continuing efforts in helioseismology will hopefully reveal the nature and strength of the magnetic field present in the solar interior and will also help in ascertaining the causes that drive the cyclic magnetic activity, and also locate the seat of the solar dynamo (cf., Choudhuri, this issue). The global and local seismology of the Sun is clearly poised to reveal its interior to a remarkably accurate detail. The uninterrupted accruing of the seismic data, can enable us to study the temporal variation of mode frequencies and amplitudes, which will indicate what changes are taking place in solar structure and dynamics. We may also learn how the Sun’s magnetic field changes with the solar activity cycle and what causes the Sun’s irradiance to vary with the sunspot cycle. 1. Cox, J. P. and Giuli, R. T., Principles of Stellar Structure, Gordon and Breach, New York, 1968. 2. Leighton, R. B., Noyes, R. W. and Simon, G. W., Astrophys. J., 1962, 135, 474–499. 3. Ulrich, R. K., Astrophys. J., 1970, 162, 993–1001. 4. Leibacher, J. and Stein, R. F., Astrophys. Lett., 1971, 7, 191– 192. 5. Deubner, F.-L., Astron. Astrophys., 1975, 44, 371–375. 6. Harvey, J. W. et al., Science, 1996, 272, 1284–1286. 7. Hill, F. et al., Science, 1996, 272, 1292–1295. 8. Scherrer, P. H. et al., Solar Phys., 1995, 162, 129–188. 9. Rhodes, E. J. Jr., Kosovichev, A. G., Schou, J., Scherrer, P. H. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS and Reiter, J., Solar Phys., 1997, 175, 287–310. 10. Rhodes, E. J. Jr., Reiter, J., Kosovichev, A. G., Schou, J. and Scherrer, P. H., Structure and Dynamics of the Interior of the Sun and Sun-like Stars (eds Korzennik, S. G. and Wilson, A.), ESA SP-418, 1998, pp. 73–82. 11. Antia, H. M. and Basu, S., Astrophys. J., 1999, 519, 400–406. 12. Goldreich, P. and Keeley, D. A., Astrophys. J., 1977, 212, 243– 251. 13. Gough, D. O. and Thompson, M. J., in Solar Interior and Atmosphere (eds Cox, A. N., Livingston, W. C. and Matthews, M.), Space Science Series, University of Arizona Press, 1991, pp. 519–561. 14. Gough, D. O. et al., Science, 1996, 272, 1296–1300. 15. Kosovichev, A. G. et al., Solar Phys., 1997, 170, 43–62. 16. Rogers, F. J. and Iglesias, C. A., Astrophys. J. Suppl., 1992, 79, 507–568. 17. Christensen-Dalsgaard, J. et al., Science, 1996, 272, 1286–1292. 18. Basu, S. and Antia, H. M., Mon. Not. R. Astron. Soc., 1995, 276, 1402–1408. 19. Christensen-Dalsgaard, J., Proffitt, C. R. and Thompson, M. J., Astrophys. J., 1993, 403, L75–L78. 20. Basu, S. and Christensen-Dalsgaard, J., Astron. Astrophys., 1997, 322, L5–L8. 21. Däppen, W., Mihalas, D., Hummer, D. G. and Mihalas, B. W., Astrophys. J., 1988, 332, 261–270. 22. Rogers, F. J., Swenson, F. J., Iglesias, C. A., Astrophys. J., 1996, 456, 902–908. 23. Elliott, J. R. and Kosovichev, A. G., Astrophys. J., 1998, 500, L199–L202. 24. Christensen-Dalsgaard, J., Gough, D. O. and Thompson, M. J., CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 Astrophys. J., 1991, 378, 413–437. 25. Basu, S., Mon. Not. R.. Astron. Soc., 1998, 298, 719–728. 26. Basu, S. and Antia, H. M., Mon. Not. R.. Astron. Soc., 1996, 287, 189–198. 27. Iglesias, C. A. and Rogers, F. J., Astrophys. J., 1996, 464, 943– 953. 28. Gough, D. O., in Progress of Seismology of the Sun and Stars, Lecture Notes in Physics (eds Osaki, Y. and Shibahashi, H.), Springer, Berlin, 1990, vol. 367, pp. 283–318. 29. Monteiro, M. J. P. F. G., Christensen-Dalsgaard, J. and Thompson, M. J., Astron. Astrophys., 1994, 283, 247–262. 30. Basu, S., Antia, H. M. and Narasimha, D., Mon. Not. R.. Astron. Soc., 1994, 267, 209–224. 31. Basu, S., Mon. Not. R.. Astron. Soc., 1997, 288, 572–584. 32. Schou, J., Kosovichev, A. G., Goode, P. R. and Dziembowski, W. A., Astrophys. J., 1997, 489, L197–L200. 33. Antia, H. M., Astron. Astrophys., 1998, 330, 336–340. 34. Gough, D. O. and Kosovichev, A. G., in Proceedings IAU Colloquium No 121, Inside the Sun (eds Berthomieu G. and Cribier M.), Kluwer, Dordrecht, 1990, pp. 327–340. 35. Takata, M. and Shibahashi, H., Astrophys. J., 1998, 504, 1035– 1050. 36. Antia, H. M. and Chitre, S. M., Astron. Astrophys., 1998, 339, 239–251. 37. Adelberger, E. C. et al., Rev. Mod. Phys., 1998, 70, 1265–1292. 38. Antia, H. M. and Chitre, S. M., Astrophys. J., 1995, 442, 434– 445. 39. Kuhn, J. R., Bush, R. I., Scheick, X. and Scherrer, P., Nature, 1998, 392, 155–157. 40. Pijpers, F. P., Mon. Not. R.. Astron. Soc., 1998, 297, L76–L80. 1459 SPECIAL SECTION: SOLAR PHYSICS Rotation of the solar interior Jørgen Christensen-Dalsgaard* and Michael J. Thompson** *Teoretisk Astrofysik Center, Danmarks Grundforskningsfond, and Institut for Fysik og Astronomi, Aarhus Universitet, DK-8000, Aarhus C, Denmark **Astronomy Unit, Queen Mary and Westfield College, University of London, UK Helioseismology has allowed us to infer the rotation in the greater part of the solar interior with high precision and resolution. The results show interesting conflicts with earlier theoretical expectations, indicating that the Sun is host to complex dynamical phenomena, so far hardly understood. This has important consequences for our ideas about the evolution of stellar rotation, as well as for models for the generation of the solar magnetic field. Here we provide an overview of our current knowledge about solar rotation, much of it obtained from observations from the SOHO spacecraft, and discuss the broader implications. SOLAR rotation has been known at least since the early seventeenth century when, with the newly invented telescope, Fabricius, Galileo and Scheiner observed the motion of sunspots across the solar disk. Indeed, at the last solar maximum 10 years ago, one of us made naked-eye observations of sunspots from the sunset walk at the TIFR, Mumbai: over several evenings the day-to-day change in position of sunspots, visible to the naked eye during the haze just before sunset, clearly showed that the Sun was rotating. It is hardly surprising that the Sun and other stars are observed to rotate. Stars are born of contracting interstellar gas clouds which share the rotation of the Galaxy. As the clouds contract, they rotate more rapidly, much as an ice skater makes herself spin around faster by pulling in her arms to her body, because she is reducing her moment of inertia while her angular momentum is conserved. Although the details of star formation within the contracting clouds are uncertain and involve mass loss and interaction with disks around the star which will transport angular momentum from one part of the cloud to another, it is plausible that newly formed stars should be spinning quite rapidly. This is indeed observed: the rotation of the stellar surface causes a broadening of the lines in the star’s spectrum, owing to the Doppler effect, and from measurements of this effect it is inferred that many young stars rotate at near the break-up speed, where the centrifugal force at the equator almost equals gravity. Stars tend to slow down when they get older. At least for *For correspondence. (e-mail: jcd@obs.aau.dk) 1460 stars of roughly solar type, the observations show that the rotation rate decreases with increasing age. The Sun’s slowdown is thought to take place through angularmomentum loss in the solar wind, magnetically coupled to the outer parts of the Sun. The extent to which the slowdown affects the deep interior of the Sun then depends on the efficiency of the coupling between the inner and outer parts. In fact, simple models of the dynamics of the solar interior tend to predict that the core of the Sun is rotating up to fifty times as rapidly as the surface. Such a rapidly rotating solar core could have serious consequences for the tests of Einstein’s theory of general relativity based on observations of planetary motion: a rapidly rotating core would flatten the Sun and hence perturb the gravitational field around it. Even a subtle effect of this nature, difficult to see directly on the Sun’s turbulent surface, might be significant. Very detailed observations have been carried out of the solar surface rotation by tracking the motion of surface features, such as sunspots and, more recently, by Dopplervelocity measurements. It was firmly established by the nineteenth century, by careful tracking of sunspots at different latitudes on the Sun’s surface, that the Sun is not rotating as a solid body: at the equator the rotation period is around 25 days, but it increases gradually towards the poles where the period is estimated to be in excess of 36 days. This differential rotation is not as surprising as it might seem: since the Sun is a sphere of gas, it is not constrained to rotate at a uniform rate. Nevertheless, the origin of the differential rotation, and how it is continued in the solar interior, are evidently interesting questions. The origin of the differential rotation is almost certainly linked to the otherwise dynamic nature of the outer parts of the Sun. The Sun’s radius is 700 Mm (i.e. 700,000 km). In the outer 200 Mm, energy is transported by convection, in rising elements of warm gas and sinking elements of colder gas: this region is called the convection zone. The convection zone can be seen directly using high-resolution observations of the solar surface, in the granulation with brighter areas of warm gas just arrived at the surface, surrounded by colder lanes of sinking gas. The gas motions also transport angular momentum, and hence provide a link between rotation in different parts of the convection zone. Also, convection is affected by rotation, which may introduce anisotropy in the angular momentum transport. Indeed, it is likely that this transport is responsible for the CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS differential rotation, although the details are far from understood. Similarly complex dynamical interactions are also found in the giant gaseous planets (Jupiter, Saturn, Uranus and Neptune) which, like the Sun, are vigorously convecting as they rotate. Here the interaction probably gives rise to the banded structures immediately visible on Jupiter, and more faintly on Saturn. Even closer to home, the Earth’s atmosphere and oceans are rotating fluid systems and exhibit, among other things, large-scale circulations and meandering jets, such as the jet stream. In all these systems, rotation plays a significant role in the observed dynamical behaviour. Helioseismic probes of the solar interior In recent years, the observation that the Sun is oscillating simultaneously in many small-amplitude global resonant modes has provided a new diagnostic of the solar interior. As discussed elsewhere in this issue (in the article by Chitre and Antia), the frequencies of these global modes depend on conditions inside the Sun, and so by measuring these frequencies we are able to make deductions about the state of the interior. This field is known as helioseismology. The observed oscillations are sometimes called five-minute oscillations, because they have periods in the vicinity of five minutes. The modes are distinguished not only by their different frequencies, but also by their different patterns on the surface of the Sun. These patterns are described by spherical harmonics (see Figure 1 for some examples) which are characterized by two integer numbers, their degree l and their azimuthal order m. As Chitre and Antia explain (see also below), different modes are sensitive to different regions of the Sun, depending on their frequency, degree and azimuthal order. By exploiting the different sensitivities of the modes, helioseismology is able to make inferences about localized conditions inside the Sun. One of the factors that affect the mode frequencies is the Sun’s rotation. The dominant effect of rotation on the oscillation frequencies is quite simple: the oscillation patterns illustrated in Figure 1 actually correspond to waves running around the equator; if the images were animated, they would essentially look like rotating beach balls. Patterns travelling in the same direction as the rotation of the Sun would appear to rotate a little faster, patterns rotating in the opposite direction a little more slowly. When Figure 1. Examples of spherical harmonic patterns for different values of the degree l and order m. Red and blue represent positive and negative regions, black represents regions where the spherical harmonic is close to zero. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1461 SPECIAL SECTION: SOLAR PHYSICS observed at a given position on the Sun, the oscillations in the former case would have slightly higher frequency, and in the latter case slightly lower frequency, than if the Sun had not been rotating. The frequency difference between these two cases therefore provides a measure of the rotation rate of the Sun. In reality other effects must be taken into account to describe the frequency shifts caused by rotation, which are often referred to as the rotational frequency splitting. The Coriolis force affects the dynamics of the oscillations and hence their frequencies, although it turns out that for the modes observed in the five-minute region this effect is modest. However, the variation of rotation rate with position in the Sun must be taken into account. Each mode feels an average rotation rate, where the average is determined by the mode’s frequency, degree and azimuthal order: the precise form of this spatial average is described by a weight function. These weight functions vary widely from mode to mode. As already indicated in Figure 1, modes with m = l are concentrated near the equator, increasingly so with increasing l, whereas modes of lower azimuthal order extend to higher latitudes. Thus modes with m = l feel only an average of the rotation near the equatorial plane, whereas modes of lower azimuthal order sense the average rotation over a wider range of latitudes. In a similar manner, as described by Chitre and Antia, the high-degree fiveminute modes (i.e. with large values of l) sense only conditions near the surface of the Sun, whereas modes of low degree feel conditions averaged over much of the solar interior. These properties can be illustrated by a few examples of weight functions (as shown in Figure 2). The observed modes include some that penetrate essentially to the solar centre, others that are trapped very near the surface, and the whole range of intermediate penetration depths, with a similar variation in latitudinal extent. Thus the observed frequency splittings provide a similarly wide range of averages of the internal rotation. It is this wealth of data which allows the determination of the detailed variation of rotation with position in the solar interior. Modes of high degree, trapped near the surface, provide measures of the rotation of the superficial layers of the Sun. Having determined that, its effect on the somewhat more deeply penetrating modes can be eliminated, leaving just a measure of rotation at slightly greater depths. In this way, information about rotation in the Sun can be ‘peeled’ layer by layer, much as one could an onion, in a way that allows us to obtain a complete image of solar internal rotation. It is fairly evident that this process gets harder, the deeper one attempts to probe, since fewer and fewer modes penetrate to the required depth; furthermore, the effect of rotation decreases because of the smaller size of the region involved. Thus the rotation of the solar core is difficult to determine. Similarly, all modes are affected by the equatorial rotation while only modes of low m extend to the vicinity of the poles, and the polar regions have relatively little effect on the oscillations, complicating the determination of the high-latitude rotation. However, as we shall see, the quality of current data is such that the rotation rate can be determined quite near the poles, at least in the outer parts of Figure 2. Weight functions determining the sensitivity of different modes to the solar internal rotation. Red indicates essentially no sensitivity, whereas green and blue show regions of successively higher sensitivity. All modes have frequencies near 2 mHz; their degree l and azimuthal order m are, from left to right: (l, m) = (5, 2), (28, 10), (28, 26), and (60, 50). 1462 CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS the convection zone. We also note an intrinsic limitation of the frequency splittings of global modes: the weight functions are symmetrical around the solar equator, as can be seen in Figure 2; thus we can infer only the similarly symmetric component of rotation. This must be kept in mind in the following, when interpreting the results. We note that this restriction can be avoided by applying local helioseismology techniques to the data: such techniques are described elsewhere in this issue by Kosovichev and Duvall. The solar internal rotation Early helioseismic data on rotational splittings provided information only about the modes with m ~ ± l; as a result, they were sensitive mainly to rotation near the equator. Observations from the Kitt Peak National Observatory, USA, showed around 1984 that there was relatively little variation of rotation with depth; in particular, there were no significant indications of a rapidly rotating core. A few years later, initial data on the dependence of the splitting on m were obtained at the Sacramento Peak and Big Bear Solar Observatories. Strikingly, they indicated that the surface latitudinal differential rotation persisted through the convection zone, whereas there was little indication of variation with latitude in the rotation beneath the convection zone. In the last few years, the amount and quality of helioseismic data on solar rotation have increased dramatically, as a result of several ground-based and space-based experiments. The LOWL instrument of the High Altitude Observatory has provided high-quality data on modes of low and intermediate degree over the past more than five years. The BiSON and IRIS networks, observing low-degree modes in Doppler velocity integrated over the solar disk, have yielded increasingly tight constraints on the rotation of the solar core, while the GONG six-station network (including a station at the Udaipur Solar Observatory) is setting a new standard for ground-based helioseismology. Finally, the SOI-MDI experiment on the SOHO spacecraft has yielded a wealth of data on modes of degree up to 300, allowing for the first time a detailed analysis of the properties of rotation in the convection zone. The results we present below are the combined knowledge that has emerged from these observational efforts. In discussing what we now know about the rotation inside the Sun, we shall start from the near-surface layers and work towards the centre. As we have already discussed, the outer 30 per cent of the Sun is convectively unstable. Before helioseismology, models predicted that the rotation inside the convection zone would organize itself on cylinders aligned with the rotation axis. Thus the rotation at depth at, say, equatorial latitudes would match the surface rotation at high latitudes, rather than the faster equatorial rotation at the surface, and so at a given CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 latitude the rotation in the convection zone would decrease with depth. Helioseismology has shown that this is not so: to a first approximation, it is more accurate to say that the rotation at a given latitude is nearly constant with depth, or to put it another way, the differential rotation seen at the surface imprints itself through the convection zone. This finding is clearly visible in Figure 3. In detail, the situation is more complicated. At low latitudes, immediately beneath the solar surface the rotation rate actually initially increases with depth. The equatorial rotation reaches a maximum at a depth of about 50 Mm (i.e. about 7 per cent of the way in from the surface to the centre of the Sun): at this point the rotation rate is about 5 per cent higher than it is at the surface. This is consistent with a variety of surface measurements of rotation. Tracking sunspots tends to give a slightly higher rotation rate than that obtained by making direct spectroscopic measurements of the velocity of the surface. Probably the reason is that the sunspots extend to some depth below the surface, and so are dragged along at a rate that is similar to the subsurface rotation a few per cent beneath the surface which helioseismology has revealed. The rotational velocity at the surface of the Sun is about 2 km sec –1 (i.e. 7000 km h –1), dropping off rather smoothly towards higher latitudes. However, it has now been found2 that superimposed on this are bands of faster and slower rotation, a few m sec –1 higher or lower than the mean flow (Figure 4). The origin of this behaviour is not understood, but it is reminiscent of the more pronounced banded flow patterns seen on Jupiter and Figure 3. Rotation of the solar envelope inferred from observations by the SOI-MDI instrument on board the SOHO satellite1 . Regions of faster rotation are red, regions of slower rotation are blue and black. The values quoted on the colour key are the frequency of the rotation (i.e. the reciprocal of the rotation period), in nano-Hertz (nHz): 300 nHz corresponds to a period of roughly 39 days, while 450 nHz corresponds to a period of about 26 days. The dashed line indicates the bottom of the convection zone. 1463 SPECIAL SECTION: SOLAR PHYSICS Saturn. Evidence for such bands had been obtained previously from direct Doppler measurements on the solar surface. However, the seismic inferences have shown that they extend to a depth probably exceeding 40 Mm beneath the surface3. Moreover, these bands migrate from high latitudes towards the equator over the solar cycle. It has been customary to represent the directly measured surface rotation rate in terms of a simple low-order expansion in sin ψ, where ψ is latitude on the Sun. This in fact quite successfully captured the observed behaviour; however, since the solar rotation axis is close to the plane of the sky, direct measurements of rotation near the poles are difficult and uncertain. Strikingly, the helioseismic results have shown a marked departure from this behaviour, at latitudes above about 60°: relative to the simple fit, the actual rotation rate decreases quite markedly there. The origin or significance of this behaviour is not yet understood. There is also evidence, hinted at in Figure 3, of a more complex behaviour of rotation at high latitudes. Some analyses have shown a ‘jet’, i.e. a localized region of more rapid rotation, at a latitude around 75° and a depth of about 35 Mm beneath the solar surface. Also, evidence has been found that the rotation rate shows substantial variations in time at high latitudes, over time scales of the order of months. It is probably fair to say that the significance of these results is still somewhat uncertain, however. Also, it should be kept in mind, as mentioned above, that the results provide an average of rotation in the northern and southern hemispheres and, evidently, an Figure 4. The evolution with time of the fine structure in the near-surface solar rotation. The time-averaged rotation rate has been subtracted from each of 11 independent inferences of rotation, for consecutive 72-day time intervals. The result is represented as a function of time (horizontal axis) and latitude (vertical axis), the colour-coding at the right gives the scale in nHz; 1.5 nHz corresponds to a speed of around 6/m sec–1 at the equator. The banded structure, apparently converging towards the equator as time goes by, should be noted. (From ref. 3.) 1464 average over the observing period of at least 2–3 months. Thus the interpretation of the inferred rotation rates in terms of the actual dynamics of the solar convection zone is not straightforward. At the base of the convection zone, a remarkable transition occurs: the variation of rotation rate with latitude disappears, so that the region beneath the convection zone rotates essentially rigidly, at a rate corresponding to the surface rate at mid-latitudes (Figure 5). The region over which the transition occurs is very narrow, no more than a few per cent of the total radius of the Sun. This layer has been called the tachocline. Why the differential rotation does not persist beneath the convection zone is not yet known, but it is possible that a large-scale weak magnetic field permeates the inner region and enforces nearly rigid rotation by dragging the gas along at a common rate5. Such a field is quite possible, as a relic from the original collapsing gas cloud from which the Sun condensed. The discovery of the tachocline, and of the form of the rotation in the convection zone, has led to an adjustment of our theories of the solar dynamo (see the article by Choudhuri in this issue). The Sun displays a roughly eleven-year cycle of sunspot activity, with the number of spots and their latitudinal distribution on the Sun varying over the cycle. Sunspots are formed where strong magnetic fields poke through the Sun’s surface, and these magnetic fields are widely believed to be generated by some kind of dynamo action in the Sun. One idea is that the dynamo action consists of two components: a twisting of the magnetic field by the motion of convective elements, and a shearing out of the field by differential rotation. Prior to the Figure 5. The inferred rotation as a function of depth inside the Sun at three solar latitudes: the equator (red), 30 degrees (orange) and 60 degrees (green). The vertical spread in the coloured bands shows the statistical uncertainty on the rotation rate (± 1 standard deviation). Note that the result becomes much more uncertain in the deep interior. The values on the vertical axis are the rotation frequency in nHz (see caption to Figure 3). The values on the horizontal axis are the fractional radius inside the Sun, and run from the centre of the Sun (r/R = 0.0) to the visible surface (r/R = 1.0). The observations used to infer the rotation were from the LOWL instrument and the BiSON network4 . CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS helioseismic findings, the simulations of rotation implied that the radial gradient of differential rotation in the convection zone could provide the second ingredient, so it was thought that the dynamo action occurred in that region. Now, however, the tachocline with its very substantial radial gradient seems a more likely location for the dynamo. Even deeper in the Sun, right down into the core where the energy-releasing nuclear reactions take place, the helioseismic results on the rotation are more uncertain due to the fact that so few of the observed five-minute modes (only the low-degree modes in fact) have any sensitivity to this region. Indeed, the results have been somewhat contradictory, some indicating rotation faster than the surface rate and others indicating rotation slower than or comparable to the rotation rate at the base of the convection zone; an example is illustrated in Figure 5. However, down to within 15 per cent of the solar radius from the centre, which is the deepest point at which present observations permit localized inferences to be made, all the modern results agree that the rotation rate is not more than a factor two different from the surface rate: thus early models which predicted that the whole of the nuclearburning core was rotating much faster are firmly ruled out. Again, this finding would be consistent with a magnetic field linking the core to the bulk of the radiative interior. Modelling solar rotation Although helioseismology has provided us with a remarkably detailed view of solar internal rotation, the theoretical understanding of the inferred behaviour is still incomplete. In the convection zone, the problem is to model the complex combined dynamics of rotation and convection, the latter occurring on scales from probably less than a few hundred kilometres to the scale of the entire convection zone and time scales from minutes to years. Viscous dissipation is estimated to occur on even smaller spatial scales, of the order 0.1 km or less. Capturing this range of scales is entirely outside the possibility of current numerical simulations; thus simplifications are required. Detailed simulations of near-surface convection, on a scale of a few Mm, have been remarkably successful in reproducing the observed properties of the granulation6, but are evidently not directly relevant to the question of rotation. Simulations of the entire convection zone are necessarily restricted to Figure 6. The results of two simulations of convection-zone rotation by Miesch, Elliott and Toomre, from the paper by Miesch9 . The two simulations use different boundary conditions and parameter values, and illustrate some of the range of possible responses of the differential rotation to the form of the convection. Also note that Simulation A has a higher resolution and includes penetration into a stable region beneath the convection zone, whereas the convective motions in Simulation B are more laminar and there is no penetration beneath the convection zone. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1465 SPECIAL SECTION: SOLAR PHYSICS rather large scales and hence cannot capture the nearsurface details. Such simulations, therefore, typically exclude the outer 30 Mm of the convection zone. Early examples of such simulations by Gilman and Glatzmaier, of fairly limited resolution, showed a tendency for rotation to organize itself on cylinders7: the rotation rate depended primarily on the distance to the rotation axis; such a behaviour is predicted for simple systems by the Taylor– Proudman theorem. The rotation rate on a given cylinder would obviously be observable where the cylinder intersected the surface; thus the observed decrease of rotation rate with increasing latitude would correspond to a similar decrease of rotation rate with depth at, say, the equator. The convection itself in these simulations was dominated by so-called banana cells – long, thin, largescale convection cells oriented in the north–south direction. The actual behaviour of rotation, shown in Figure 3, is obviously very different from these simulation results. The overall variation of rotation within the convection zone is evidently predominantly with latitude, with little variation in the radial direction except in the tachocline. Given the necessary simplifications of the calculations, their failure to model solar rotation is perhaps not surprising. In particular, the effects of smaller-scale turbulence (beneath the smallest scale resolved in the simulation) are typically represented as some form of viscosity; it was suggested by Gough, and later by others, that the effect of rotation on the small-scale motion might render this turbulent viscosity non-isotropic, with important effects on the transport of angular momentum within the convection zone. In fact, simple models of convection- zone dynamics, with parametrized anisotropic viscosity, have had some success in reproducing the helioseismically inferred rotation rate. Recent advances in computing power have led to improved numerical simulations8, which come closer to representing turbulent convective flow regimes such as exist in the Sun’s convection zone. Figure 6 shows results from two such simulations by Miesch, Elliott and Toomre. The simulations can yield a range of differential rotation profiles, depending on the conditions imposed at the top and bottom boundaries of the simulation region, and on the parameter values adopted for the problem. Since it is not obvious what are the most appropriate boundary 1466 conditions and parameter values to choose, it is necessary to explore various possibilities and study the different responses. Simulation B has rotation contours at mid-latitudes which are nearly radial, as in the Sun (compare Figure 3), but the contrast in rotation rate between low and high latitudes is not as great as is observed in the Sun (about 70 nHz, rather than 130 nHz). In case A, the latitudinal variation of the Sun’s rotation is better reproduced, but the mid-latitude contours do not look quite as similar to those in Figure 3. Nonetheless, these results are encouraging indications that we may be close to reproducing theoretically the gross features of the solar rotation inferred by helioseismology. There is still, though, much work ahead, both observational and theoretical, in getting a detailed understanding of the Sun’s rotation and with that, we hope, a better understanding of the solar activity cycle and of large-scale rotating fluid systems on planets and stars. 1. Schou, J. et al., Astrophys. J., 1998, 505, 390–417. 2. Schou, J. and the SOI Internal Rotation Team, in Proceedings IAU Symposium 185: New Eyes to See Inside the Sun and Stars (eds Deubner, F.-L., Christensen-Dalsgaard, J. and Kurtz, D. W.), Kluwer, Dordrecht, 1998, pp. 141–148. 3. Toomre, J., Christensen-Dalsgaard, J., Howe, R., Larsen, R. M., Schou, J. and Thompson, M. J., Solar Phys., 1999, 308, 405– 414. 4. Chaplin, W. J. et al., Mon. Not. R. Astron. Soc., 1999 (in press). 5. Gough, D. O. and McIntyre, M. E., Nature, 1998, 394, 755– 757. 6. Stein, R. F. and Nordlund, Å., Astrophys. J., 1989, 342, L95– L98. 7. Gilman, P. A. and Miller, J., Astrophys. J. Suppl., 1986, 61, 585–608. 8. Elliott, J. R., Miesch, M. S., Toomre, J., Clune, T. and Glatzmaier, G. A.,. in Structure and Dynamics of the Interior of the Sun and Sun-like Stars; Proc. SOHO 6/GONG 98 Workshop (eds Korzennik, S. G. and Wilson, A.), European Space Agency, Noordwijk, The Netherlands, 1998, pp. 765–770. 9. Miesch, M. S., Solar Phys., 1999 (in press). ACKNOWLEDGEMENTS. We are grateful to Dr M. Miesch for providing Figure 6, and to Dr R. Howe for help with other Figure. We acknowledge the financial support of the UK Particle Physics and Astronomy Research Council, and the Danish National Research Foundation through its establishment of the Theoretical Astrophysics Center. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Solar tomography A. G. Kosovichev*,† and T. L. Duvall Jr** *W.W. Hansen Experimental Physics Laboratory, Stanford University, CA 95305-4085, USA **Laboratory for Astronomy and Solar Physics, NASA Goddard Space Flight Center Greenbelt, MD 20771, USA The solar tomography (or time-distance helioseismology) is a new promising method for probing 3-D structures and flows beneath the solar surface, which is potentially important for studying the birth of active regions in the Sun’s interior and for understanding the relation between the internal dynamics of the active regions, and the chromospheric and coronal activity. In this method, the time for waves to travel along sub-surface ray paths is determined from the temporal cross correlation of signals at two separated surface points. By measuring the times for many pairs of points from Dopplergrams, covering the visible hemisphere, a tremendous quantity of information about the state of the solar interior is derived. As an example, we present the results on the internal structures of supergranulation, meridional circulation, active regions and sunspots. An active region which emerged on the solar disk in January 1998, was studied from SOHO/MDI for nine days, both before and after its emergence at the surface. The results show a complicated structure of the emerging region in the interior, and suggest that the emerging flux ropes travel very quickly through the depth range of our observations. Method of solar tomography solar surface1 T Ψ (τ, ∆) = ∫ f (t , r1 ) f * ( t + τ , r2 ) dt , 0 (1) where ∆ is the horizontal distance between the points with coordinates r1 and r2, τ is the delay time, and T is the total time of the observations. Because of the stochastic nature of excitation of the oscillations, function Ψ must be averaged over some areas on the solar surface to achieve a signal-to-noise ratio sufficient for measuring travel times τ. The oscillation signal, f(t, r), is usually the Doppler velocity or intensity. A typical cross-covariance function shown in Figure 2 displays several sets of ridges which correspond to the first, second, and third bounces of packets of acousticwave packets from the surface. The origin of the multiple bounces is illustrated in Figure 1. Waves originated at point A may reach point B directly (solid curve), or after one-bounce at point C (dashed curve), or after two-bounces (dotted curve), and so on. Because the sound speed is greater in the deeper layers, the direct waves arrive first, followed by the second-bounce and third-bounce waves. The cross-covariance function represents a solar ‘seismogram’. Figure 3 shows the cross-covariance signal as a Solar acoustic waves are excited by turbulent convection near the solar surface and propagate through the interior with the speed of sound. Because the sound speed increases with depth, the waves are refracted and reappear on the surface at some distance away from the source. The wave propagation is illustrated in Figure 1. The waves excited at point A will reappear at the surface points B, C, D, E, F, and others after propagating along the ray paths indicated by curves. The basic idea of solar tomography is to measure the acoustic travel time between different points on the solar surface, and then to use these measurements for inferring variations of the structure and flow velocities in the interior along the wave paths connecting the surface points. This idea is similar to the Earth’s seismology. However, unlike in Earth, the solar waves are generated stochastically by numerous acoustic sources in the subsurface layer of turbulent convection. Therefore, the wave travel-time is determined from the cross-covariance function, Ψ(τ, ∆), of the oscillation signal, f(t, r), between different points on the † For correspondence (e-mail: sasha@guake.stanford.edu) CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 Figure 1. A cross-section diagram through the solar interior, illustrating the wave propagation inside the Sun. 1467 SPECIAL SECTION: SOLAR PHYSICS function of time for a distance of 30 degrees. It consists of three wave packets corresponding to the first, second, and third bounces. Ideally, the seismogram should be inverted to infer the structure and flows using a wave theory. However, in practice, as in terrestrial seismology2, different approximations are employed, the most simple and powerful of which is the geometrical acoustic (ray) approximation. Generally, the observed solar oscillation signal corresponds to displacement or pressure perturbation, and can be represented in terms of normal modes, or standing waves. Therefore, the cross-covariance function can be expressed in terms of normal modes, and then represented as a superposition of traveling wave packets3. An example of the theoretical cross-covariance function of p-modes of the standard solar model is shown in Figure 4. This model reproduces the observational cross-covariance function very well in the observed range of distances, from 0 to 90 degrees. The theoretical model was calculated for larger distances, including points on the far side of the Sun, which are not accessible for observation. A backward propagating ridge originating from the second-bounce ridge at 180 degrees is a geometrical effect due to the choice of the range of the angular distance from 0 to 180 degrees. This is illustrated by the green curve in Figure 1. The waves reaching point F after the reflection at point D propagate more than 180 degrees, but are considered as propagating the distance AF which is less than 180 degrees. In the theoretical diagram (Figure 4) one can notice a weak backward ridge between 30 and 70 degrees and at 120 min. This ridge is due to reflection from the solar core. However, it has not been detected in observations. By grouping the modes in narrow ranges of the angular phase velocity, ν = ωnl/L, where L = (l(l+ 1))1/2, and applying the method of stationary phase, the cross- Figure 3. The observed cross-covariance signal as a function of time at the distance of 30 degrees. Figure 2. The observational cross-covariance function as a function of distance on the solar surface, ∆, and the delay time, τ. The lowest set of ridges (first-bounce) corresponds to waves propagated to the distance, ∆, without additional reflections from the solar surface. The middle ridge (second-bounce) is produced by the waves arriving to the same distance after one reflection from the surface, and the upper ridge (third-bounce) results from the waves arriving after two bounces from the surface. The backward ridge associated with the second-bounce ridge is due to the choice of the angular distance range from 1468 Figure 4. The model cross-covariance function calculated from the p-modes theoretical eigenfunctions. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS covariance function can be approximately represented in the form3 2 δω2 ∆ ∆ Ψ (τ, ∆ ) ∝ ∑ cos ω0 τ − exp − τ − , ν 4 u δv (2) where δv is a narrow interval of the phase speed, u ≡ (∂ω/∂k h) is the horizontal component of the group velocity, k h = L/R is the angular component of the wave vector, R is the solar radius, ω0 is the central frequency of a Gaussian frequency filter applied to the data, and δω is the characteristic width of this filter. Therefore, the phase- and group travel-times are measured by fitting individual terms of eq. (2) to the observed cross-covariance function using a least-squares technique. This technique measures both phase (∆/v ) and group (∆/u) travel-time of the p-mode wave packets. The previous time-distance measurements provided either the group-time4, or the phase times5. It was found that the noise level in the phase-time measurements was substantially lower than in the grouptime measurements. Therefore, we use the phase times in this paper. We also employ the geometrical acoustic (ray) approximation to relate the measured-phase times to the internal properties of the Sun. More precisely, the variations of the local travel-times at different points on the surface relative to the travel-times averaged over the observed area are used to infer variations of the internal structure and flow velocities using a perturbation theory. In the ray approximation, the travel-times are sensitive only to the perturbations along the ray paths given by the Hamilton equations. The variations of the travel-time obey Fermat’s Principle6 δτ = 1 δ k d r, ω ∫Γ (3) where δk is the perturbation of the wave vector due to the structural inhomogeneities and flows along the unperturbed ray path, Γ. Using the dispersion relation for acous- tic waves in the convection zone, the travel-time variations can be expressed in terms of the sound speed, magnetic field strength and flow velocity3. The effects of flows and structural perturbations are separated from each other by taking the difference and the mean of the reciprocal travel-times δτdiff ≈ −2 ∫ Γ δτmean ≈ − ∫ Γ ( nU ) c2 ds ; δc S ds , c (4) (5) where c is the adiabatic sound speed, n is a unit vector tangent to the ray, S = k/ωis the phase slowness. Magnetic field causes anisotropy of the mean travel-times, which allows us to separate, in principle, the magnetic effects from the variations of the sound speed (or temperature). So far, only a combined effect of the magnetic fields and temperature variations has been measured reliably. Onedimensional tests by Kosovichev and Duvall3 and twodimensional numerical simulations by Jensen et al.7 have shown that eqs (4) and (5) provide a reasonable approximation to the travel-time variations. The development of a more accurate theory for the travel-times, based on the Born approximation is currently under way. Typically, we measure times for acoustic waves to travel between points on the solar surface and surrounding quadrants symmetrical relative to the North, South, East and West directions. In each quadrant, the travel-times are averaged over narrow ranges of travel distance ∆. Then, the times for northward-directed waves are subtracted from the times for south-directed waves to yield the time, NS τ diff which predominantly measures north–south motions. EW Similarly, the time differences, τ diff , between westwardand eastward-directed waves yields a measure of eastward oi motion. The time, τ diff , between outward- and inwarddirected waves, averaged over the full annuli, is mainly sensitive to vertical motion and the horizontal divergence. Figure 5. The regions of ray propagation (colour areas) as a function of depth, z, and the radial distance, ∆, from a point on the surface. The rays are also averaged over circular regions on the surface, forming three-dimensional figures of revolution. The dashed lines show the inversion grid. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1469 SPECIAL SECTION: SOLAR PHYSICS The time, τmean , which measures sound speed and magnetic perturbations, is also averaged over the full annuli (see refs 3 and 8). The next step is to determine the variations of the sound speed and flow velocity from the observed travel-times using eqs (4) and (5). It is assumed that the convective structures and flows do not change during the observations and can be represented by a discrete model. In this model, the 3-D region of wave propagation is divided into rectangular blocks. The perturbations of the sound speed, and the three components of the flow velocity are approximated by the linear functions of coordinates within each block, e.g. for the flow velocity U ( x, y, z ) = | x − xi | i +1 − x i ∑ U ijk 1 − x | y − yj | 1 − y j +1 − y j | z − zk | 1 − z k +1 − z k , flows. However, vertical flows in the deep layers are not resolved because of the predominantly horizontal propagation of the rays in these layers. The vertical velocities are also systematically underestimated by 10–20% in the upper layers. Similarly, the sound speed variations are underestimated in the bottom layers. These limitations of the solar tomography should be taken into account in interpretation of the inversion results. Inversion results Helioseismic tomography has been successfully used to infer local properties of large-scale zonal and meridional flows11, convective flows and structures (refs 3 and 8), structure and dynamics of active regions12 and flows around sunspots 5. Here we present some results of tomo- (6) where xi, yj, zk are the coordinates of the rectangular grid, Uijk are the values of the velocity in the grid points, and xi ≤ x ≤ xi + 1, yj ≤ y ≤ yj + 1, and zk ≤ z ≤ zk + 1. A part of the x – z grid is shown in Figure 5 together with the ray system used in inversion. The travel-time measured at a point on the solar surface is the result of the cumulative effects of the perturbations in each of the traversed rays of the 3-D ray systems. Figure 5 shows, in the ray approximation, the sensitivity to subsurface location for a certain point on the surface. This pattern is then translated for different surface points in the observed area, so that overall the travel-times are sensitive to all subsurface points in the depth range 0–20 Mm, in this example. We average the equations over the ray systems corresponding to the different radial distance intervals of the data, using approximately the same number of ray paths as in the observational procedure. As a result, we obtain two systems of linear equations that relate the data to the sound speed variation and to the flow velocity, e.g. for the velocity field, δτ diff; λµν = ∑ A ijk λµν ⋅ U ijk , (7) ijk where vector-matrix A maps the structure properties into the observed travel-time variations, and indices λ and µ label the location of the central point of a ray system on the surface, and index ν labels the annuli. These equations are solved by a regularized least-squares technique using the LSQR algorithm9. Jensen et al.10 suggested to speed up the inversion by doing most of the calculation in the Fourier domain. The results of test inversions of Kosovichev and Duvall3 demonstrate a very accurate reconstruction of sound speed variations and the horizontal components of subsurface 1470 Figure 6. The horizontal flow velocity field (arrows) and the sound speed perturbation (red colour shows positive perturbations, and blue colour shows negative perturbations) at the depths of (a) 1.4 Mm and (b) 5.0 Mm as inferred from the SOHO/MDI high-resolution data of 27 January 1996. The arrows at the South–North axis indicate the location of the vertical cut in the East–West direction, which is shown in Figure 7. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS graphic inversion for large-scale convective cells (supergranulation), meridional flow, sunspot and an emerging active region. Quiet-Sun convection The data used were for 8.5 h on 27 January 1996 from the high-resolution mode of the MDI instrument. The results of inversion of these data are shown in Figure 6. We have found that, in the upper layers, 2–3 Mm deep, the horizontal flow is organized in supergranular cells, with outflows from the centre of the supergranules. The characteristic size of the cells is 20–30 Mm. Comparing with MDI magnetograms, it was found that the cell boundaries coincide with the areas of enhanced magnetic field. These results are consistent with the observations of supergranulation on the solar surface13. However, in the layers deeper than 5 Mm, we do not see the supergranulation pattern. This suggests that supergranulation is only 5 Mm deep. An alternative interpretation suggesting a depth of 8 Mm was presented by Duvall14. The vertical flows (Figure 7) correlate with the supergranular pattern in the upper layers. Typically, there are upflows in the ‘hotter’ areas and downflows in the ‘colder’ areas. In the hotter areas however, the sound speed is higher than the average. at low latitudes to higher latitudes and, therefore, contribute to the cyclic polar-field reversal. The meridional flows in the solar interior were detected by the time-distance method. Figure 8 shows the differences between the travel-times of acoustic waves propagating poleward and equatorward at different latitudes λ. These travel-time differences correspond to the mean meridional flow averaged over the penetration depth of the acoustic waves, which was 4–24 Mm in the measurements. By using eq. (4) Giles et al.11 estimated that the maximum mean speed of the flow is ~ 20 m s –1. They have also found that the flow velocity is almost constant over the observed range of depth. Tomography of sunspots and active regions An important problem of astrophysics is understanding the mechanisms of solar activity. The solar tomography provides a tool for studying the birth and evolution of active regions and complexes of solar activity. In Figure 9, we show the results for the emerging active region observed in January 1998. This was a high- Meridional circulation Meridional flows from the equator towards the north and south poles have been observed on the solar surface in direct Doppler-shift measurements15. The MDI observations by Giles et al.11 have provided the first evidence that such flows persist to great depths, and, thus, possibly play an important role in the 11-year solar cycle. The poleward flow can transport the magnetic remnants of sunspots generated Figure 7. The vertical-flow field (arrows) and the sound speed perturbation at the North–South position indicated by arrows in Figure 6. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 Figure 8. The average travel-time difference (south minus north) as a function of latitude, λ, for surface separation of pairs of points in the range 12–73 Mm. The individual points are shown (squares) and the 1σ errors (vertical lines). The solid curve is the best-fit 2parameter model. The velocity scale on the right axis, in which 12.1 m/s flow corresponds to a 1 s time difference, is obtained from eq. (3) (ref. 11). 1471 SPECIAL SECTION: SOLAR PHYSICS latitude region of the new solar cycle which was started in 1997. Figure 9 shows the distribution of the wave speed variations in a vertical cross-section in the region of the emerging flux and in a horizontal plane at a depth of 18 Mm, for three 2-h intervals, a, at 17:00 UT, 11 January 1998; b, 3:00 UT, 12 January 1998; and c, 5:00 UT, 12 January 1998. The perturbations of the magnetosonic speed shown in this figure are associated with the magnetic field and temperature variations in the emerging magnetic ropes. The positive variations are shown in red, and the negative variations are shown in blue. Figure 9 a shows no significant variations in the region of the emergence, which is at the middle of the vertical plane. The MDI magnetogram shown at the top indicates only very weak magnetic field above this region. Figure 9 b shows a positive perturbation associated with the emerging region. The strongest perturbation in this panel is at the bottom of the observed region. During the next 2 h (Figure 9 c), the perturbation is propagated to the top of the box. From these data, we conclude that the emerging flux propagated through the characteristic depth of 10 Mm during 2 h. This gives an estimate of the speed of emergence ≈ 1.3 km/s. This speed is somewhat higher than the speed predicted by theories of emerging flux. The typical amplitude of the sound speed variation in the perturbation is about 0.5 km/s. This may correspond to a magnetic field strength of 500 G at the top of the box, or a temperature variation of 800 K. After the emergence we observed the gradual increase of the perturbation in the subsurface layers with the formation of sunspots. The observed development of the active region suggests that the sunspots were formed as result of the concentration of magnetic flux close to the surface. Figure 10 is an example of the internal structure of a large sunspot observed on 17 January 1998. An image of the spot taken in the continuum is shown at the top. The sound speed perturbations in the spot are much stronger than in the emerging flux, and reach more than 3 km/s. It is interesting that beneath the spot the perturbation is negative in the subsurface layers and becomes positive in the deeper interior. This data also shows the connection to the spot of a small pore which is on the left side of the spot. The negative perturbations beneath the spot are, probably due to the lower temperature. However, the effects of temperature and magnetic field have not been separated in these inversions. Separating these effects is an important problem of solar tomography. Figure 9. The sound speed perturbation in the emerging active region (a) on 11 January 1998, 17:00 UT; (b) 12 January 1998, 3:00; and (c) 12 January 1998, 5:00. The horizontal size of the box is 415 Mm, the vertical size is 18 Mm. The panels on the top are MDI magnetogram showing the surface magnetic field of positive (red) and negative (blue) polarities. The perturbations of the sound speed range from – 1.6 to 1.3 km/s. The positive variations are shown in red, and the negative ones in blue. 1472 CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Figure 10. The sound-speed perturbation in a sunspot region observed on 17 January 1998. The horizontal size of the box is 158 Mm, the depth is 24 Mm. mechanisms of solar activity. Conclusion Solar tomography, or time-distance helioseismology, provides unique information about three-dimensional structures and flows associated with magnetic field and turbulent convection in the solar interior. This method is at the very beginning of its development. In this paper, we have reviewed some basic principles of this technique, based on the geometrical ray approximation, and presented some initial inversion results. Developing wave-form solar tomography is one of the most challenging problems of helioseismology. Using time-distance seismology, we have been able to measure the structure of supergranulation flows and detect an active region before it appeared on the surface. The inversion results also have shown interesting dynamics of supergranulation, meridional circulation, emerging active regions and the formation of sunspots in the upper convection zone. Further studies of the Sun’s interior by the time-distance seismology will shed light on the CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1. Duvall, T. L., Jr., Jefferies, S. M., Harvey, J. W. and Pomerantz, M. A., Nature, 1993, 362, 430–432. 2. Aki, K. and Richards, P., Quantitative Seismology. Theory and Methods, Freeman, San Francisco, 1980. 3. Kosovichev, A. G. and Duvall, T. L. Jr., in SCORe’96: Solar Convection and Oscillations and their Relationship (eds Pijpers, F. P., Christensen-Dalsgaard, J. and Rosenthal, C. S.), Kluwer Academic Publishers, 1997, pp. 241–260. 4. Jefferies, S. M., Osaki, Y., Shibahashi, H. et al., Astrophys. J., 1994, 434, 795–800. 5. Duvall, T. L. Jr., D’Silva, S., Jefferies, S. M., Harvey, J. W. and Schou, J., Nature, 1996, 379, 235–237. 6. Gough, D. O., in Astrophysical Fluid Dynamics (eds Zahn, J.-P. and Zinn-Justin, J.), Elsevier Science Publ., 1993, p. 339. 7. Jensen, J. M., Jacobsen, B. H. and Christensen-Dalsgaard, J., 1999, preprint. 8. Duvall, T. L. Jr., Kosovichev, A. G., Scherrer, P. H. et al., Solar Phys., 1997, 170, 63–73. 9. Paige, C. C. and Saunders, M. A., ACM Trans. Math. Software, 1982, 8, 43–71. 10. Jensen, J. M., Jacobsen, B. H. and Christensen-Dalsgaard, J., Proceedings of the Interdisciplinary Inversion Workshop 5, Aarhus, 1997, pp. 57–67. 1473 SPECIAL SECTION: SOLAR PHYSICS 11. Giles, P. M., Duvall, T. L. Jr. and Scherrer, P. H., Nature, 1997, 390, 52–54. 12. Kosovichev, A. G., Astrophys. J., 1996, 461, L55–L57. 13. Title, A. M., Tarbell, T. D., Topka, K. P., Ferguson, S. H. and Shine, R. A., Astrophys. J., 1989, 336, 475–494. 14. Duvall, T. L. Jr., Proceedings of the SOHO-6/GONG-98 Workshop, ESA, 1998. ACKNOWLEDGEMENTS. The authors acknowledge many years of effort by the engineering and support staff of the MDI development team at the Lockheed Palo Alto Research Laboratory (now Lockheed–Martin Advanced Technology Center) and the SOI development team at Stanford University. SOHO is a project of international cooperation between ESA and NASA. This research is supported by the SOI-MDI NASA contract NAG5-3077 at Stanford University. 15. Duvall, T. L. Jr., Solar Phys., 1979, 63, 3–15. 1474 CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1475 SPECIAL SECTION: SOLAR PHYSICS The solar dynamo Arnab Rai Choudhuri Department of Physics, Indian Institute of Science, Bangalore 560 012, India It is believed that the magnetic field of the Sun is produced by the dynamo process, which involves nonlinear interactions between the solar plasma and the magnetic field. After summarizing the main characteristics of solar magnetic fields, the basic ideas of dynamo theory are presented. Then an appraisal is made of the current status of solar dynamo theory. 1. Introduction IN elementary textbooks on stellar structure, a star is usually modelled as a spherically symmetric, non-rotating, nonmagnetic object. It is mainly the magnetic field which makes our Sun much more intriguing than such a textbook star. Several other reviews in this special section should convince the reader of this. It comes, therefore, as no surprise that one of the central problems in solar physics is to understand the origin of the Sun’s magnetic field. The solar dynamo theory attempts to address this problem. The basic idea of this theory is that the solar magnetic fields are generated and maintained by complicated nonlinear interactions between the solar plasma and magnetic fields. As we shall see in this review, there are still many difficulties with this theory and we are still far from having a completely satisfactory explanation of why the Sun’s magnetic field behaves the way it does. However, no alternate theory of the origin of solar magnetism has so far been able to explain even a fraction of what dynamo theory has explained. Some of us, therefore, are still struggling to put the solar dynamo theory on firmer footing, with the fond hope that we are probably approximately on the correct path. The aim of this special section is to make the readers of Current Science aware of the present status of solar physics. The solar dynamo theory is a fairly technical subject. It is next to impossible to write a review that will provide a comprehensive introduction to this subject for an average reader of Current Science and, at the time, survey the research frontiers. Still a partial attempt is made here at this next-to-impossible task of presenting the subject in a way which should be understandable – if not to a general reader of Current Science – at least to a reader with some familiarity in physics and fluid mechanics. It is left to the readers to judge if the author has failed completely or only moderately. Needless to say, no attempt is made at a e-mail: arnab@physics.iisc.ernet.in CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 complete coverage of the fundamentals. After summarizing the relevant observations in the following section, we write just enough about the basics in the next two sections to give a rough idea of what is going on. Then in the last three sections we discuss some of the important issues from current research frontiers. Dynamo theory is based on the principles of magnetohydrodynamics (MHD), in which hydrodynamics equations are combined with Maxwell’s electrodynamics equations. Comprehensive introductions to MHD can be found in the books by Alfvén and Fälthammar1, Cowling2, Parker3, Priest4, and Choudhuri5. Some books devoted exclusively to dynamo theory are by Moffatt6, Krause and Rädler7, and Zeldovich et al.8. We also refer to the review articles on the solar dynamo by Ruzmaikin9, Gilman10, Hoyng11, Brandenburg and Tuominen12, and Schmitt13. 2. Relevant observations In 1908 Hale14 discovered the first evidence of Zeeman effect in sunspot spectra and made the momentous announcement that sunspots are regions of strong magnetic fields. This is the first time that somebody found conclusive evidence of large-scale magnetic fields outside the Earth’s environment. The typical magnetic field of a large sunspot is about 3000 G. Even before it was realized that sunspots are seats of solar magnetism, several persons have been studying the occurences of sunspots. Schwabe15 noted that the number of sunspots seen on the solar surface increases and decreases with a period of about 11 years. Now we believe that the Sun has a cycle with twice that period, i.e. 22 years. Since the Sun’s magnetic field changes its direction after 11 years, it takes 22 years for the magnetic field to come back to its initial configuration. Carrington16 found that sunspots seemed to appear at lower and lower latitudes with the progress of the solar cycle. In other words, most of the sunspots in the early phase of a solar cycle are seen between 30° and 40°. As the cycle advances, new sunspots are found at increasingly lower latitudes. Then a fresh halfcycle begins with sunspots appearing again at high latitudes. Individual sunspots live from a few days to a few weeks. After finding magnetic fields in sunspots, Hale and his coworkers17 made another significant discovery. They found that often two large sunspots are seen side by side and they invariably have opposite polarities. The line joining the centres of such a bipolar sunspot pair is usually 1475 SPECIAL SECTION: SOLAR PHYSICS nearly parallel to the solar equator. Hale’s coworker Joy, however, noted that there is a systematic tilt of this line with respect to the equator and that this tilt increases with latitude17. This result is usually known as Joy’s Law. It was also noted17 that the sunspot pairs have opposite polarities in the two hemispheres. In other words, if the left sunspot in the northern hemisphere has negative polarity, then the left sunspot in the southern hemisphere has positive polarity. This is clearly seen in Figure 1, which is a magnetic map of the Sun’s disk obtained with a magnetogram. The regions of positive and negative polarities are shown in white and black respectively. The polarities of the bipolar sunspots in any hemisphere get reversed from one half-cycle of 11 years to the next half-cycle. After the development of the magnetograph by Babcock and Babcock18, it became possible to study the much weaker magnetic field near the poles of the Sun. This magnetic field is of the order of 10 G and reverses its direction at the time of solar maximum19 (i.e. when the number of sunspots seen on the solar surface is maximum). This shows that this weak, diffuse field of the Sun is in some way coupled to the much stronger magnetic field of the sunspots and is a part of the same solar cycle. Lowresolution magnetograms show the evidence of weak magnetic field even in lower latitudes. The true nature of this field is not very clear. It was found20 that the magnetic field on the solar surface outside sunspots often exists in the form of fibril flux tubes of diameter of the order of 300 km with field strength of about 2000 G (large sunspots have sizes larger than 10,000 km). One Figure 1. A magnetogram image of the full solar disk. The regions with positive and negative magnetic polarities are respectively shown in white and black, with grey indicating regions where the magnetic field is weak. Courtesy: K. Harvey. 1476 is not completely sure if the field found in the lowresolution magnetograms is truly a diffuse field or a smearing out of the contributions made by fibril flux tubes. Keeping this caveat in mind, we should refer to the field outside sunspots as seen in magnetograms as the ‘diffuse’ field. It was found21 that there were large unipolar matches of this diffuse field on the solar surface which migrated towards the pole. Even when averaged over longitude, one finds predominantly one polarity in a belt of latitude which drifts polewards22,23. The reversal of polar field presumably takes place when sufficient field of opposite polarity has been brought near the poles. Figure 2 (taken from Dikpati and Choudhuri24) shows the distribution of both sunspots and the weak, diffuse field in a plot of latitude vs. time. The colour shades indicate values of longitude-averaged diffuse field, whereas the latitudes where sunspots were seen at a particular time are marked by vertical black lines. The sunspot distribution in a timelatitude plot is often referred to as a butterfly diagram, since the pattern (the vertical black lines in Figure 2) reminds one of a butterfly. Such butterfly diagrams were first plotted by Maunder25. Historically, most of the dynamo models concentrated on explaining the distribution of Figure 2. Colour-shades showing the latitude-time distribution of longitudinally averaged weak, diffuse magnetic field (B is in Gauss) with a ‘butterfly diagram’ of sunspots superimposed on it during the interval from May 1976 to December 1985. Reproduced from Dikpati and Choudhuri24 . CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS sunspots and ignored the diffuse field. Only during the last few years, it has been realized that the diffuse fields give us important clues about the dynamo process and they should be included in a full self-consistent theory. The aim of such a theory should be to explain diagrams like Figure 2 (i.e. not just the butterfly diagram). We have provided above a summary of the various regular features in the Sun’s activity cycle. One finds lots of irregularities and fluctuations superposed on the underlying regular behaviour, as can be seen in Figure 2. These irregularities are more clearly visible in Figure 3, where the number of sunspots seen on the solar surface is plotted against time. Galileo was one of the first persons in Europe to study sunspots at the beginning of the 17th century. After Galileo’s work, sunspots were almost not seen for nearly a century26! It may be noted that all the observations discussed above pertain to the Sun’s surface. We have no direct information about the magnetic field underneath the Sun’s surface. The new science of helioseismology, however, has provided us lots of information about the velocity field underneath the solar surface. For an account of this subject, the readers may turn to the reviews by Chitre and Antia, and by Christensen-Dalsgaard and Thompson. We shall have occasions to refer to some of the helioseismic findings in our discussion later. It is to be noted that heat is transported by convection in the outer layers of the Sun from about 0.7 R¤ to R¤ (where R¤ is the solar radius). This region is called the convection zone, within which the plasma is in a turbulent state. The job of a theorist now is to construct a detailed model of the physical processes in this turbulent plasma such that all the surface observations of magnetic fields are properly explained – a fairly daunting problem, of which the full solution is still a distant dream. 3. Some basic magnetohydrodynamics considerations The velocity field v and the magnetic field B in a plasma (regarded as a continuum) interact with each other according to the following MHD equations: ∂v 1 B 2 (B ⋅ ∇ ) B + + ( v ⋅ ∇ ) v = ∇ p + + g + ν∇ 2 v , (1) ∂t ρ 8π 4πρ ∂B 2 = ∇ × ( v × B) + λ∇ B . ∂t Here ρ is density, p is pressure, g is gravitational field, ν is kinematic viscosity, and λ= c2 4πσ CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 (3) is magnetic diffusivity (σ is electrical conductivity). Equation (1) is essentially the Navier–Stokes equation, to which magnetic forces have been added. It is clear from eq. (1) that the magnetic field has two effects: (i) it gives rise to an additional pressure B2/8π; and (ii) the other magnetic term (B · ∇)B/4πρ is of the nature of a tension along magnetic field lines. Equation (2) is known as the induction equation and is the key equation in MHD. It has the same form as the vorticity equation in ordinary hydrodynamics (see, for example, § 4.2 and § 5.2 of Choudhuri5). If V, B and L are the typical values of velocity, magnetic field and length scale, then the two terms on the RHS of eq. (2) are of order VB/L and λB/L2. The ratio of these two terms is a dimensionless number, known as the magnetic Reynolds number, given by Rm = Figure 3. The number of sunspots seen in a year plotted against the year for the period 1610–1975. The original figure is due to John A. Eddy. Reproduced from Moffatt6 . (2) VB /L λB/L2 = VL . λ (4) Since Rm goes as L, it is expected to be much larger in astrophysical situations than it is in the laboratory. In fact, usually one finds that Rm >> 1 in astrophysical systems and Rm << 1 in laboratory-size objects. Hence the behaviours of magnetic fields are very different in laboratory plasmas and astrophysical plasmas. For example, it is not possible to have laboratory analogues of the selfsustaining magnetic fields of the Earth or the Sun. If Rm >> 1 in an astrophysical system, then the diffusion term in eq. (2) is negligible compared to the term preceding it. In ordinary hydrodynamics, when the viscous dissipation term in the vorticity equation is neglected, we are led to the famous 1477 SPECIAL SECTION: SOLAR PHYSICS Kelvin’s theorem of vorticity conservation (see, for example, § 4.6 of Choudhuri5). Exactly similarly, when the diffusion term in eq. (2) is neglected, it can be shown that the magnetic field is frozen in the plasma and moves with it. This result was first recognized by Alfvén27 and is often referred to as Alfvén’s Theorem of Flux-Freezing. It is known that the Sun does not rotate like a solid body. The angular velocity at the equator is about 20% faster than that at the poles. Because of the flux freezing, this differential rotation would stretch out any magnetic field line in the toroidal direction (i.e. the φ direction with respect to the Sun’s rotation axis). This is indicated in Figure 4. We, therefore, expect that the magnetic field inside the Sun may be predominantly in the toroidal direction. We have already mentioned in § 2 that energy is transported by convection in the layers underneath the Sun’s surface. To understand why the magnetic field remains concentrated in structures like sunspots instead of spreading out more evenly, we need to study the interaction of the magnetic field with the convection in the plasma. This subject is known as magnetoconvection. The linear theory of convection in the presence of a vertical magnetic field was studied by Chandrasekhar28. The nonlinear evolution of the system, however, can only be found from numerical simulations pioneered by Weiss29. It was found that space gets separated into two kinds of regions. In certain regions, magnetic field is excluded and vigorous convection takes place. In other regions, magnetic field gets concentrated and the tension of magnetic field lines suppresses convection in those regions. Sunspots are presumably such regions where magnetic field is piled up by surrounding convection. Since heat transport is inhibited there due to the suppression of convection, sunspots look darker than the surrounding regions. Although we have no direct information about the state of the magnetic field under the Sun’s surface, it is expected that the interactions with convection would keep the magnetic field concentrated in bundles of field lines throughout the solar convection zone. Such a concentrated bundle of magnetic field lines is called a flux tube. In the a a b b Figure 4. The production of a strong toroidal magnetic field underneath Figure 5. the Magnetic Sun’s surface. buoyancy a.ofAn a flux initial tube. poloidal a. A nearly field horizontal line. b. A sketch flux tubeofunder the field the solar linesurface. after itb.has Thebeen flux tube stretched after by its upper the faster part rotation has risennear through the equatorial the solar surface. region. 1478 regions of strong differential rotation, therefore, we may have flux tubes aligned in the toroidal direction. If a part of such a flux tube rises up and pierces the solar surface as shown in Figure 5 b, we expect to have two sunspots with opposite polarities at the same latitudes. But how can a configuration like Figure 5 b arise? The answer to this question was provided by Parker30 through his idea of magnetic buoyancy. We have seen in eq. (1) that a pressure B2/8π is associated with a magnetic field. If p in and p out are the gas pressures inside and outside a flux tube, then we need to have p out = pin + B2 8π (5) to maintain pressure balance across the surface of a flux tube. Hence, p in ≤ p out, (6) which often, though not always, implies that the density inside the flux tube is less than the surrounding density. If this happens in a part of the flux tube, then that part becomes buoyant and rises against the gravitational field to produce the configuration of Figure 5 b starting from Figure 5 a. A look at Figure 4 now ought to convince the reader that the sub-surface toroidal field in the two hemispheres should have opposite polarity. If this toroidal field rises due to magnetic buoyancy to produce the bipolar sunspot pairs, we expect the bipolar sunspots to have opposite polarities in the two hemispheres as seen in Figure 1. We thus see that combining the ideas of flux freezing, magnetoconvection and magnetic buoyancy, we can understand many aspects of the bipolar sunspot pairs. We now turn our attention to the central problem – the dynamo generation of the magnetic field. 4. The turbulent dynamo and mean field MHD We now address the question whether it is possible for motions inside the plasma to sustain a magnetic field. Ideally, one would like to solve eqs (1) and (2) to understand how velocity and magnetic fields interact with each other. Solving these two equations simultaneously in any non-trivial situation is an extremely challenging job. In the early years of dynamo research, one would typically assume a velocity field to be given and then solve eq. (2) to find if this velocity field would sustain a magnetic field. This problem is known as the kinematic dynamo problem. One of the first important steps was a negative theorem due to Cowling31, which established that an axisymmetric solution is not possible for the kinematic dynamo problem. One is, therefore, forced to look for more complicated, non-axisymmetric solutions. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS A major breakthrough occurred in 1955 when Parker32 realized that the turbulent motions inside the solar convection zone (which are by nature non-axisymmetric) may be able to sustain the magnetic field. We have indicated in Figure 4 how a magnetic field line in the poloidal plane may be stretched by the differential rotation to produce a toroidal component. Parker32 pointed out that the uprising hot plasma blobs in the convection zone would rotate as they rise because of the Coriolis force of solar rotation (just like cyclones in the Earth’s atmosphere) and such helically moving plasma blobs would twist the toroidal field shown in Figure 6 a to produce magnetic loops in the poloidal plane as shown in Figure 6 b. Keeping in mind that the toroidal field has opposite directions in the two hemispheres and helical motions of convective turbulence should also have opposite helicities in the two hemispheres, we conclude that the poloidal loops in both hemispheres should have the same sense as indicated in Figure 6 c. Although we are in a high magnetic Reynolds number situation and the magnetic field is nearly frozen in the plasma, there is some diffusion (especially due to turbulent mixing) and the poloidal loops in Figure 6 c should eventually coalesce to give the largescale poloidal field as sketched by the broken line in Figure 6 c. Figure 7 captures the basic idea of Parker’s turbulent dynamo. The poloidal and toroidal components of the magnetic field feed each other through a closed loop. The poloidal component is stretched by differential rotation to produce the toroidal component. On the other hand, the helical turbulence acting on the toroidal component gives back the poloidal component. Parker32 developed a heuristic mathematical formalism based on these ideas and showed by mathematical analysis that these ideas worked. However, a more systemic mathematical formulation of these ideas had to await a few years, when Steenbeck, Krause and Rädler33 developed what is known as mean field MHD. Some of the basic ideas of mean field MHD are summarized below. Since we have to deal with a turbulent situation, let us split both the velocity field and the magnetic field into average and fluctuating parts, i.e. v = v + v ′, B = B + B ′. is known as the mean e.m.f. and is the crucial term for dynamo action. This term can be perturbatively evaluated by a scheme known as the first-order smoothing approximation (see, for example, § 16.5 of Choudhuri5). If the turbulence is isotropic, then this approximation scheme leads to ε = αB − β ∇ × B, (10) where 1 α = − v ′ ⋅ (∇ × v ′) τ , 3 (11) and β= 1 v ′ ⋅ v ′τ . 3 (12) Here τ is the correlation time of turbulence. On substituting (10) in eq. (8), we get ∂B = ∇ × ( v × B) + ∇ × (αB) + ( λ + β)∇ 2 B . ∂t (13) It should be clear from this that β is the turbulent diffusion. This is usually much larger than the molecular diffusion λ so that λ can be neglected in eq. (13). It follows from eq. (11) that α is a measure of average helical motion in the fluid. It is this coefficient which describes the production of the poloidal component from the toroidal component by helical turbulence. This term would go to zero if turbulence has no net average helicity (which would happen in a non-rotating a b (7) Here the overline indicates the average and the prime indicates the departure from the average. On substituting eq. (7) in the induction eq. (2) and averaging term by term, we obtain ∂B = ∇ × ( v × B ) + ∇ × ε + λ∇ 2 B ; ∂t (8) on remembering that v ′ = B′ = 0 .Here, ε = v ′ × B′ CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 c (9) Figure 6. 7. Different Schematicstages representation of the dynamo of Parker’s process. idea See text of the for explanation. turbulent dynamo. 1479 SPECIAL SECTION: SOLAR PHYSICS frame). Equation (13) is known as the dynamo equation and has to be solved to understand the generation of magnetic field by the dynamo process. A variant of this equation was first derived by rather intuitive arguments in the classic paper of Parker32. The mean field MHD developed by Steenbeck, Krause and Rädler33 put this equation on a firmer footing. In the kinematic dynamo approach, one has to specify a velocity field v and then solve eq. (13). Using spherical polar coordinates with respect to the rotation axis of the Sun, we can write v = Ω (r , θ ) r sin θ eφ + v P , (14) where Ω(r, θ) is the angular velocity in the interior of the Sun and v P is some possible average flow in the poloidal plane. Until a few years ago, almost all the calculations of the kinematic dynamo problem were done by taking v P = 0. If this is the case, then one has to specify some reasonable Ω(r, θ) and α(r, θ) before proceeding to solve the dynamo eq. (13). In the 1970s almost an industry grew up presenting solutions of the dynamo equation for different specifications of Ω and α. The first pioneering solution in rectangular geometry was obtained by Parker32 himself. He showed that periodic and propagating wave solutions of the dynamo equation are possible. Presumably this offers an explanation for the solar cycle. Sunspots migrate from higher to lower latitudes with the solar cycle because sunspots are produced (by magnetic buoyancy) where the crest of the propagating dynamo wave lies. Parker32 found that the parameters α and Ω have to satisfy the following condition in the northern hemisphere to make the dynamo wave propagate in the equatorward direction (so as to explain the butterfly diagram of sunspots): α dΩ ≤ 0. dr (15) Steenbeck and Krause34 were the first to solve the dynamo equation in a spherical geometry appropriate for the Sun and produced the first theoretical butterfly diagram of the distribution of sunspots in time-latitude. Then many dynamo solutions were worked out by Roberts35, Köhler36, Yoshimura37, Stix38 and others. One might have felt complacent about the varieties of butterfly diagrams produced by these authors. However, it has to be admitted that many basic physics questions remained unanswered. Since nothing was known at that time about the conditions in the interior of the Sun, different authors were choosing different α and Ω subject only to the condition (15), and thereby were trying to fit the observational data better. Eventually it appeared that it was becoming a game in which you could get solutions according to your wishes by tuning your free parameters suitably. Further progress in solar dynamo theory became possible only by 1480 asking fundamental questions about the basic physics in the interior of the Sun, rather than by blindly solving the dynamo equation. These efforts will be described in the next section. It may be noted that all the authors of this period focussed their attention on explaining the equatorward propagation of sunspots, by assuming that sunspots were produced in the regions where the toroidal component had the peak value. No serious attempt was made to connect the behaviour of the weak, diffuse magnetic field with the dynamo process or to explain the poleward migration of this field, although Köhler36 and Yoshimura37 presented some models that show a polar branch, i.e. a region near the poles where the dynamo wave propagates poleward. 5. Dynamo in the overshoot layer? Where does the solar dynamo work? Since one needs convective turbulence to drive the dynamo, it used to be tacitly assumed in the early 1970s that the dynamo works in the solar convection zone and the different researchers of that period used to take α(r, θ) non-zero in certain regions of the convection zone. This approach had to be questioned when Parker39 started looking at the effect of magnetic buoyancy on the solar dynamo. Magnetic buoyancy is particularly destabilizing in the interior of the convection zone, where convective instability and magnetic buoyancy reinforce each other. On the other hand, if a region is stable against convection, then magnetic buoyancy can be partially suppressed there (see, for example, § 8.8 of Parker3). Calculations of buoyant rise by Parker39 showed that any magnetic field in the convection zone would be removed from there by magnetic buoyancy fairly quickly. Hence it is difficult to make the dynamo work in the convection zone, since the magnetic field has to be stored in the dynamo region for a sufficient time to allow for dynamo amplification. It is expected that there is a thin overshoot layer (probably with a thickness of the order of 104 km) just below the bottom of the convection zone. This is a layer which is convectively stable according to a local stability analysis, but convective motions are induced there due to convective plumes from the overlying unstable layers overshooting and penetrating there. Several authors (Spiegel and Weiss40, van Ballegooijen41) pointed out that this layer is a suitable location for the operation of the dynamo. Although there would be enough turbulent motions in this layer to drive the dynamo, magnetic buoyancy would be suppressed by the stable temperature gradient there. This idea turned out to be a really prophetic theoretical guess, since helioseismology observations a few years later indeed discovered a region of strong differential rotation at the bottom of the solar convection zone. See the review by Christensen-Dalsgaard and Thompson in this issue on this subject. So it is certainly expected that a strong toroidal magnetic field CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS should be generated just below the bottom of the convection zone due to this strong differential rotation. It may be noted that there have been other ideas as well for suppressing magnetic buoyancy at the bottom of the convection zone. Parker42 suggested ‘thermal shadows’, whereas van Ballegooijen and Choudhuri43 showed that an equatorward meridional circulation at the base of the convection zone can help in suppressing magnetic buoyancy there. For about a decade starting from the mid-1980s, most researchers in this field believed that the whole dynamo process in the Sun, as summarized in Figure 7, takes place in the overshoot layer. Properties of such a dynamo operating in the overshoot layer were studied by DeLuca and Gilman44, Gilman et al.45, and Choudhuri46. If the dynamo operates in the overshoot layer, some new questions arise. Previously when the solar dynamo was supposed to work in the convection zone, the sunspots seen on the solar surface could be regarded as direct signatures of the dynamo process. One could assume that sunspots appeared wherever the dynamo produced strong toroidal fields just underneath the surface. On the other hand, if the dynamo works at the bottom of the convection zone, the whole depth of the convection zone separates the region where the magnetic fields are generated and the solar surface where sunspots are seen. In order to understand the relation between the solar dynamo and sunspots, one then has to study how the magnetic fields generated at the bottom of the convection zone rise through the convection zone to produce sunspots. The best way to study this is to treat it as an initial-value problem. First an initial configuration with some magnetic flux at the bottom of the convection zone is specified, and then its subsequent evolution is studied numerically. The evolution depends on the strength of magnetic buoyancy, which is in turn determined by the value of the magnetic field. If the dynamo is driven by turbulence, one would expect an equipartition of energy between the dynamogenerated magnetic field and the fluid kinetic energy, i.e. B2 1 2 ≈ ρv . 8π 2 (16) This suggests B ≈ 104 G on the basis of standard models of convection. Because of the strong differential rotation, we expect the magnetic field at the bottom of the convection zone to be mainly in the toroidal direction. One, therefore, has to take a toroidal magnetic flux tube going around the CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 rotation axis as the initial configuration. The evolution of such magnetic flux tubes due to magnetic buoyancy was first studied by Choudhuri and Gilman47 and Choudhuri48. It was found that the Coriolis force due to the Sun’s rotation plays a much more important role in this problem than what anybody suspected before. If the initial magnetic field is taken to have a strength around 104 G, the flux tubes move parallel to the rotation axis and emerge at very high latitudes rather than at latitudes where sunspots are seen. Only if the initial magnetic field is taken as strong as 105 G, magnetic buoyancy is strong enough to overpower the Coriolis force and the magnetic flux tubes can rise radially to emerge at low latitudes. D’Silva and Choudhuri49 extended these calculations to look at the tilts of emerging bipolar regions at the surface. Figure 8 taken from their paper shows the observational tilt vs. latitude plot of bipolar sunspots (i.e. Joy’s law) along with the theoretical plots obtained by assuming different values of the initial magnetic field. It is clearly seen that theory fits observations only if the initial magnetic field is about 105 G. Apart from providing the first quantitative explanation of Joy’s law nearly three-quarters of a century after its discovery, these calculations put the first stringent limit on the value of the toroidal magnetic field at the bottom of the convection zone. Several other groups50–53 soon performed similar calculations and confirmed the result. The evidence is now mounting that the magnetic field at the bottom of the convection zone is indeed much stronger than the equipartition value given by eq. (16) (see Schüssler54 for a review of this topic). If the magnetic field is much stronger than the equipartition value, it would be impossible for the helical turbulence (entering the mathematical theory through the α term defined in eq. (11)) to twist the magnetic field lines. The dynamo process, as envisaged in Figure 7, is therefore, not possible. We need a very different type of dynamo model. Schmitt55 and Feriz-Mas et al.56 proposed that the buoyant instability of the strong magnetic field itself may lead to a magnetic configuration which was previously thought to be created by helical turbulence. Parker57 suggested an ‘interface dynamo’ in which the helical turbulence acts in a region above the bottom of the convection zone. This idea has been further explored by Charbonneau and MacGregor58. In the next section, we discuss what we regard as the most promising approach to build a model of the solar dynamo that can account for the very strong toroidal magnetic field at the bottom of the convection zone. 1481 SPECIAL SECTION: SOLAR PHYSICS 6. The Babcock–Leighton approach and hybrid models We saw in § 4 and § 5 that one of the crucial ingredients in turbulent dynamo theory is the role of helical turbulence in generating the poloidal component from the toroidal component, which is mathematically modelled through mean field MHD. This approach to the dynamo problem will be called the Parker–Steenbeck–Krause–Rädler or the PSKR approach. In this approach, the dynamo is supposed to operate within a region where convective turbulence exists and no attention is paid to phenomena taking place at the solar surface. Babcock59 and Leighton60 in the 1960s developed a somewhat different approach, which we call the Babcock–Leighton or the BL approach. Even in this approach, the toroidal magnetic field is believed to be produced by the differential rotation of the Sun. For the production of the poloidal component, however, a totally different scenario is invoked. The strong toroidal component leads to bipolar sunspots due to magnetic buoyancy. We have noted that these bipolar sunspots have a tilt with respect to latitudinal lines (i.e. Joy’s law). Therefore, when these bipolar sunspots eventually decay, the magnetic flux spreads around in such a way that the flux at the higher latitude has more contribution from the polarity of the sunspot which was at the higher latitude. In this way, a poloidal component arises. Compared to the PSKR approach, the BL approach was heuristic and semi-qualitative. The mean field MHD is, no doubt, based on some assumptions and approximations, and it is not clear whether these hold in the conditions prevailing in the Sun’s interior. However, for an ideal system satisfying these assumptions and approximations, the mean field MHD is a rigorous mathematical theory. No quantitative mathematical theory of comparable sophis- Figure 8. Plots of sin (tilt) against sin (latitude) theoretically obtained for different initial values of magnetic field indicated in kG. The observational data indicated by the straight line fits the theoretical curve for initial magnetic field 100 kG (i.e. 105 G). Reproduced from D’Silva and Choudhuri49 . 1482 tication was developed for the BL approach. Nearly all the self-consistent dynamo calculations in the 1970s and 1980s, therefore, followed the PSKR approach. The BL approach was developed further by a group in NRL23,61–64 who were studying the spread of magnetic flux from the decay of sunspots. There was growing evidence that there is a general meridional flow with an amplitude of about a few m s –1 near the Sun’s surface proceeding from the equator to the pole65. The poloidal magnetic field produced from the decay of tilted bipolar sunspots is carried poleward by this meridional circulation. As we have already pointed out in § 2, the weak diffuse magnetic field on the solar surface migrates towards the pole, in contrast to sunspots which migrate equatorward. One presumably has to identify the weak diffuse field as the poloidal component of the Sun’s magnetic field, whereas sunspots form from the much stronger toroidal component. The main aim of the NRL group was to model the evolution of the weak diffuse field, assuming that this was entirely coming from the decay of bipolar sunspots. No attempt was made to address the full dynamo problem. They even made the drastically simple assumption that the magnetic field is a scalar residing on the solar surface and the appropriate partial differential equation was solved only on this two-dimensional surface. Dikpati and Choudhuri66,24, and Choudhuri and Dikpati67 attempted to make a vectorial model of the evolution of the weak diffuse field and to connect it to the dynamo problem. Since the meridional flow at the surface is poleward, there must be an equatorward flow in the lower regions of the convection zone, rising near the equator. If the dynamo operated at the base of the convection zone, then, in accordance with the ideas prevalent a few years ago, the poloidal component produced by this dynamo would be brought to the surface by the meridional circulation. This can be an additional source of the weak diffuse field at the surface, apart from the contributions coming from the decay of sunspots. Figure 9 from Choudhuri and Dikpati67 shows a theoretical time-latitude distribution of the weak diffuse field on the surface, obtained by assuming a dynamo wave at the bottom of the convection zone as given. In other words, to produce this figure – which should be compared with the observational Figure 2 – the dynamo problem was not solved self-consistently. The aim now should be to develop a self-consistent model of the dynamo, which should be able to explain the behaviours of both the sunspots (i.e. the toroidal component) and the weak diffuse field (i.e. the poloidal component). With helioseismology establishing the existence of strong differential rotation at the base of the convection zone, there is little doubt that the toroidal magnetic field is produced there and has a magnitude of the order of 105 G, a result pinned down by the simulations of buoyant flux rise. It will, however, be impossible for helical turbulence to twist Figure 9. Amagnetic theoreticalfields, time-latitude of the weak, such strong and it distribution seems improbable that diffuse magnetic field on the solar surface, with ‘half-butterfly the generation of the poloidal diagrams’ obtained from a running dynamo wave assumed given at the of the fromturbulence Choudhuri–and fieldbasefrom theconvection toroidal zone. fieldReproduced by helical as 68 Dikpati . CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS envisaged in the PSKR approach – takes place at the base of the convection zone. For the generation of the poloidal field, we then invoke the BL idea that it is produced by the decay of bipolar sunspots on the surface. The meridional circulation can then carry this field poleward, to be eventually brought to the bottom of the convection zone where it is stretched by the differential rotation to produce the strong toroidal field. If a mean field formulation is made of the process of poloidal field generation near the surface by invoking an α coefficient concentrated near the solar surface, this hybrid model of the dynamo will incorporate the best features of both the PSKR and the BL approaches. On the one hand, detailed quantitative calculations will be possible, as in the PSKR approach. On the other hand, the surface phenomena emphasized in the BL approach, are integrated in the dynamo problem. The meridional circulation plays an important role in this hybrid model so that a suitable form of in eq. (14) is to be specified. It is hoped that this hybrid model will account for both the equatorward migration of the strong toroidal field at the base of the convection zone and the poleward drift of the poloidal field at the surface. This hybrid model has one other attractive feature. Researchers in 1970s built kinematic models of the solar dynamo by arbitrarily specifying α(r, θ) and Ω(r, θ). These were regarded as free parameters to be tuned suitably so as to give solutions with desired characteristics. In the present hybrid models, these important ingredients to the dynamo process are directly based on observations. Helioseismology has given us Ω(r, θ), with its shear concentrated at the base of the convection zone. The α coefficient also arises out of the observed decay of bipolar sunspots on the surface. Previously it used to be even debated whether α is positive or negative. Researchers used to fudge α such that the inequality (15) was satisfied. The direction of tilt of bipolar sunspots on the surface, however, clearly indicates that α arising out of their decay has to be positive in the northern hemisphere: a point made by Stix38 long ago. Once these key ingredients are fixed directly by observation, we no longer have the freedom to fudge them according to our wishes, which researchers in 1970s could do. This leads to one problem. Helioseismology shows that dΩ/dr is positive in lower latitudes where sunspots are seen. If α is also positive in the northern hemisphere, then clearly inequality (15) is not satisfied and dynamo waves are expected to propagate poleward! The first calculations on the hybrid model were reported by Choudhuri et al.68. Dynamo waves were indeed found to propagate poleward, if meridional circulation was switched off. The toroidal field and the poloidal field are respectively produced in layers near the base of the convection zone and near the solar surface. When meridional circulation is switched off, any field can reach from one layer to the other layer by diffusion with a time scale of L2/β, where L is the separation between the layers (i.e. the thickness of the convection zone). If the meridional CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 flow has a typical velocity of the order V, it takes time V/L for the meridional circulation to carry some quantity between the two layers. When the time scale V/L is shorter than the diffusion time scale L2/β, the problem is dominated by meridional circulation and Choudhuri et al.68 found that the strong toroidal component at the bottom of the convection zone actually propagates equatorward, overriding the inequality (15). Thus the inequality (15), which was regarded as sacrosanct for four decades since Parker32 obtained it, is found not to hold in the presence of a meridional circulation having a time scale shorter than diffusion time, thereby opening up the possibility of constructing realistic hybrid models of the solar dynamo. Further calculations on hybrid models have been reported in a series of papers69–73. It should be emphasized that all these studies are still of rather exploratory nature. They demonstrate the viability of the hybrid models and study their different characteristics. We are, however, still far from building a sufficiently realistic model, putting in all the details, that would account for the observational data presented in Figure 2. Achieving this should be our goal now. 7. Miscellaneous ill-understood issues Since this is a review in a special section on solar physics, we have primarily discussed those aspects of kinematic dynamo models which directly pertain to the matching of theory with observational data. It should, however, be kept in mind that many fundamental issues of dynamo theory are still very ill-understood. Until we have a better understanding of these issues, the kinematic models can, at best, be considered superficial attempts at a very deep physics problem. The reader may look up the IAU Symposium volume on The Cosmic Dynamo74 for several articles dealing with these fundamental issues. Here we make only very brief comments on some of these issues. We have seen in § 4 that the turbulent dynamo theory is developed by averaging over turbulent fluctuations. The existence of magnetic flux concentrations clearly indicates that the fluctuations are much larger than the average values (often by orders of magnitude). Does a mean field theory make sense in such a situation? Can we trust the perturbative procedures like the first-order smoothing approximation? Hoyng75 raised some questions regarding the interpretation of the averaged quantities. The dynamo eq. (13) admits of several possible modes in spherical geometry: the preferred mode seems to be the mode with dipole symmetry, wherein the toroidal component is oppositely directed in the two hemispheres. This mode approximately corresponds to the observational data. However, Stenflo and Vogel76 pointed out that one hemisphere of the Sun often has more sunspots than the other, indicating that there may be a superposition with higher modes having different symmetry. Analysing the 1483 SPECIAL SECTION: SOLAR PHYSICS statistics of sunspot data for several decades, Gokhale and Javaraiah77 claimed to have found evidence for multiple modes. If the fluctuations are so large, there is no reason why a particular mode should be very stable, or why higher modes should not be excited. The interference of modes with different symmetry was theoretically studied by Brandenburg et al.78 employing a nonlinear dynamo model. Since the toroidal magnetic field is far stronger than the equipartition value, it is certainly not justified to assume that the magnetic fields do not back-react on the flow. One should therefore ideally solve eqs (1) and (2) simultaneously, instead of proceeding with kinematic models. Since this is a fairly difficult job even by the standard of today’s computers, attempts are made to include the backreaction of the magnetic field within kinematic models instead of going to fully dynamic models. One easy way to incorporate the back-reaction in the dynamo eq. (13) is to make the crucial quantity αdecrease with the magnetic field, following some prescription like: α= α0 1 + ( B /B0 )2 . (17) The effect of such α-quenching on the dynamo process has been extensively studied79–82. When the velocity field v is specified, the dynamo eq. (13) is a linear equation for the magnetic field provided B we assume the various coefficients in the equation to be independent of B. If α is quenched by B, in accordance with eq. (17), we have a nonlinear problem. One important question is whether the irregularities of the solar cycle, as seen in Figure 3, can be explained with the help of nonlinear models. It seems that the nonlinearity introduced through eq. (17) cannot cause such chaotic behaviour. Since a sudden increase in the amplitude of magnetic field would diminish the dynamo activity by reducing α and thereby pull down the amplitude again (a decrease in the amplitude would do the opposite), the α-quenching mechanism tends to lock the system in a stable mode once the system relaxes to it. In fact, Krause and Meinel83 argued that nonlinearities must be what makes one particular mode of the dynamo so stable. Only by introducing more complicated kinds of nonlinearity (with suppression of differential rotation) in some highly truncated dynamo models, Weiss et al.84 were able to find the evidence of chaos. Jennings and Weiss 85 presented a study of symmetry-breakings and bifurcations in a nonlinear dynamo model. Since α-quenching of the form (17) cannot explain the irregularities of the solar cycle, Choudhuri86 explored the effect of stochastic fluctuations on the mean equations and obtained some solutions resembling Figure 3. Several subsequent papers87–89 explored this possibility further. Finally we comment on the efforts in building fully dynamic models by solving both eqs (1) and (2) simultaneously. This is a highly complicated nonlinear problem and can only be tackled numerically. Gilman90 and Glatz1484 maier91 presented very ambitious numerical calculations in which convection, differential rotation and dynamo process were all calculated together from the basic MHD equations. These calculations, however, gave results which do not agree with observational data. For example, angular velocity was found to be constant on cylinders, whereas helioseismology found it to be constant on cones. If various diffusivities were set such that the surface rotation pattern was matched, the dynamo waves propagated from the equator to the pole. The codes of Gilman90 and Glatzmaier91 naturally had finite grids, and the physics at the sub-grid scales was modelled by introducing various eddy diffusivities. Probably the physics at sub-grid scales is more subtle and the details of it are crucially important in determining the behaviour of the dynamo. This is generally believed to be the reason why these massive codes did not produce agreement with observations. The subsequent approach in numerical modelling has been to do dynamic calculations over cubes which correspond to small regions of the Sun, rather than trying to build models for the whole Sun. Brandenburg et al.92 and Nordlund et al.93 have followed this approach. 8. Conclusion It seems that the solar magnetic fields are generated and maintained by the dynamo process. There is only a small minority of solar physicists who would disagree with this point of view. It is, however, not easy to build a sufficiently detailed and realistic model of the dynamo process to account for all the different aspects of solar magnetism. The 1970s happened to be a period of optimism in dynamo research when various researchers were producing butterfly diagrams by choosing different forms of α(r, θ) and Ω(r, θ). It was felt that further research would narrow down the parameter space and establish a standard model of the solar dynamo. As we discussed above, this did not happen. In 1993 Schmitt13 wrote in his review on the solar dynamo: ‘The original hope that detailed observational and theoretical information would yield better results of the dynamo did not prove true, on the contrary, they raised difficulties instead’. Today, a few years afterwards, we can perhaps have a less pessimistic outlook. The hybrid models, which are closely linked to observations and which combine together some of the best ideas that came out of dynamo research in the last few decades, certainly do look promising. Although sufficiently detailed models have not yet been worked out, we hope that we are close to building kinematic models which are much more realistic and sophisticated than the kinematic models of the 1970s. Perhaps other researchers may regard this point of view as a reflection of this author’s personal prejudice. Only time will tell if this prejudice is justified. Finally we end by cautioning the reader that this article should not be regarded as a comprehensive review of the CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS solar dynamo problem. Since this article is primarily aimed not at the experts but at more general readers, we have emphasized those aspects of kinematic models which have direct relevance to observations. A very incomplete discussion of various fundamental issues is presented in the § 7. There is no doubt that kinematic models can never fully satisfy us. The ultimate challenge is to build fully dynamic models starting from the basic equations, and then to explain both the fluid flow patterns and magnetic patterns in the interior of the Sun in a grand scheme. As we have pointed out in § 7, the modern computers still seem inadequate for handling this problem. The solar dynamo problem will certainly remain alive for years to come. 1. Alfvén, H., and Fälthammar, C.-G., Cosmical Electrodynamics, Oxford University Press, 1963, 2nd edn. 2. Cowling, T. G., Magnetohydrodynamics, Adam Hilger, 1976, 2nd edn. 3. Parker, E. N., Cosmical Magnetic Fields, Oxford University Press, 1979. 4. Priest, E. R., Solar Magnetohydrodynamics, D. Reidel, 1982. 5. Choudhuri, A. R., The Physics of Fluids and Plasmas: An Introduction for Astrophysicists, Cambridge University Press, 1998. 6. Moffatt, H. K., Magnetic Field Generation in Electrically Conducting Fluids, Cambridge University Press, 1978. 7. Krause, F. and Rädler, K.-H., Mean-Field Magnetohydrodynamics and Dynamo Theory, Pergamon, 1980. 8. Zeldovich, Ya B., Ruzmaikin, A. A. and Sokoloff, D. D., Magnetic Fields in Astrophysics, Gordon and Breach, 1983. 9. Ruzmaikin, A. A., Solar Phys., 1985, 100, 125. 10. Gilman, P. A., in Physics of the Sun (ed. Sturrock, P. A.), D. Reidel, 1986, vol. 1, p. 95. 11. Hoyng, P., in Basic Plasma Processes on the Sun, IAU-Symp. No. 142 (eds. Priest, E. R. and Krishan, V.), Kluwer, Dordrecht, 1990, p. 45. 12. Brandenburg, A. and Tuominen, I., in The Sun and Cool Stars: Activity, Magnetism, Dynamos (eds. Tuominen, I., Moss, D. and Rüdiger, G.), Lecture Notes in Physics, 1990, 380, 223. 13. Schmitt, D., in The Cosmic Dynamo (eds. Krause, F., Rädler, K.-H. and Rüdiger, G.), IAU-Symp. No. 157, Kluwer, 1993, p. 1. 14. Hale, G. E., Astrophys. J., 1908, 28, 315. 15. Schwabe, S. H., Astron. Nachr., 1844, 21, 233. 16. Carrington, R. C., Mon. Not. R. Astron. Soc., 1858, 19, 1. 17. Hale, G. E., Ellerman, F., Nicholson, S. B. and Joy, A. H., Astrophys. J., 1919, 49, 153. 18. Babcock, H. W. and Babcock, H. D., Astrophys. J., 1955, 121, 349. 19. Babcock, H. D., Astrophys. J., 1959, 130, 364. 20. Stenflo, J. O., Solar Phys., 1973, 32, 41. 21. Bumba, V. and Howard, R., Astrophys. J., 1965, 141, 1502. 22. Howard, R. and LaBonte, B. J., Solar Phys., 1981, 74, 131. 23. Wang, Y.-M., Nash, A. G. and Sheeley, N. R., Astrophys. J., 1989, 347, 529. 24. Dikpati, M. and Choudhuri, A. R., Solar Phys., 1995, 161, 9. 25. Maunder, E. W., Mon. Not. R. Astron. Soc., 1904, 64, 747. 26. Eddy, J. A., Science, 1976, 192, 1189. 27. Alfvén, H., Ark. f. Mat. Astr. o. Fysik, 1942, B29. 28. Chandrasekhar, S., Philos. Mag., 1952, 43, 501. 29. Weiss, N. O., J. Fluid Mech., 1981, 108, 247. 30. Parker, E. N., Astrophys. J., 1955, 121, 491. 31. Cowling, T. G., Mon. Not. R. Astron. Soc., 1934, 94, 39. 32. Parker, E. N., Astrophys. J., 1955, 122, 293. 33. Steenbeck, M., Krause, F. and Rädler, K.-H., Z. Naturforsch., 1966, 21a, 1285. 34. Steenbeck, M. and Krause, F., Astron. Nachr., 1969, 291, 49. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. Roberts, P. H., Philos. Trans. R. Soc. London, 1972, 272, 663. Köhler, H., Astron. Astrophys., 1973, 25, 467. Yoshimura, H., Astrophys. J. Suppl., 1975, 29, 467. Stix, M., Astron. Astrophys., 1976, 47, 243. Parker, E. N., Astrophys. J., 1975, 198, 205. Spiegel, E. A. and Weiss, N. O., Nature, 1980, 287, 616. van Ballegooijen, A. A., Astron. Astrophys., 1982, 113, 99. Parker, E. N., Astrophys. J., 1987, 312, 868. van Ballegooijen, A. A. and Choudhuri, A. R., Astrophys. J., 1988, 333, 965. DeLuca, E. E. and Gilman, P. A., Geophys. Astrophys. Fluid Dyn., 1986, 37, 85. Gilman, P. A., Morrow, C. A. and DeLuca, E. E., Astrophys. J., 1989, 338, 528. Choudhuri, A. R., Astrophys. J., 1990, 355, 733. Choudhuri, A. R. and Gilman, P. A., Astrophys. J., 1987, 316, 788. Choudhuri, A. R., Solar Phys., 1989, 123, 217. D’Silva, S. and Choudhuri, A. R., Astron. Astrophys., 1993, 272, 621. Fan, Y., Fisher, G. H. and DeLuca, E. E., Astrophys. J., 1993, 405, 390. Schüssler, M., Caligari, P., Ferriz-Mas, A. and Moreno-Insertis, F., Astron. Astrophys., 1994, 281, L69. Caligari, P., Moreno-Insertis, F. and Schüssler, M., Astrophys. J., 1995, 441, 886. Longscope, D. and Fisher, G. H., Astrophys. J., 1996, 458, 380. Schüssler, M., in The Cosmic Dynamo (eds. Krause, F., Rädler, K.-H. and Rüdiger, G.), IAU-Symp. No. 157, Kluwer, Dordrecht, 1993, p. 27. Schmitt, D., Astron. Astrophys., 1987, 174, 281. Ferriz-Mas, A., Schmitt, D. and Schüssler, M., Astron. Astrophys., 1994, 289, 949. Parker, E. N., Astrophys. J., 1993, 408, 707. Charbonneu, P. and MacGregor, K. B., Astrophys. J., 1997, 486, 502. Babcock, H. W., Astrophys. J., 1961, 133, 572. Leighton, R. B., Astrophys. J., 1969, 156, 1. DeVore, C. R., Sheeley, N. R. and Boris, J. P., Solar Phys., 1984, 92, 1. Sheeley, N. R., DeVore, C. R. and Boris, J. P., Solar Phys., 1985, 98, 219. Sheeley, N. R., Wang, Y.-M. and Harvey, J. W., Solar Phys., 1989, 119, 323. Wang, Y.-M., Nash, A. G. and Sheeley, N. R., Science, 1989, 245, 712. Duvall, T. L., Solar Phys., 1979, 63, 3; LaBonte, B. J. and Howard, R., Solar Phys., 1982, 80, 361; Ulrich et al., Solar Phys., 1988, 117, 291. Dikpati, M. and Choudhuri, A. R., Astron. Astrophys., 1994, 291, 975. Choudhuri, A. R. and Dikpati, M., Solar Phys., 1999, 184, 61. Choudhuri, A. R., Schüssler, M. and Dikpati, M., Astron. Astrophys., 1995, 303, L29. Durney, B. R., Solar Phys., 1995, 160, 213. Durney, B. R., Solar Phys., 1996, 166, 231. Durney, B. R., Astrophys. J., 1997, 486, 1065. Dikpati, M. and Charbonneau, P., Astrophys. J., 1999 (in press). Nandi, D. and Choudhuri, A. R., Astrophys. J., 1999 (submitted). Krause, F., Rädler, K.-H. and Rüdiger, G. (eds.), The Cosmic Dynamo, IAU Symp. No. 157, Kluwer, Dordrecht, 1993. Hoyng, P., Astrophys. J., 1988, 332, 857. Stenflo, J. O. and Vogel, M., Nature, 1986, 319, 285. Gokhale, M. H. and Javaraiah, J., Mon. Not. R. Astron. Soc., 1990, 243, 241. Brandenburg, A., Krause, F., Meinel, R., Moss, D. and Tuominen, I., Astron. Astrophys., 1989, 213, 411. Stix, M., Astron. Astrophys., 1972, 20, 9. Jepps, S. A., J. Fluid Mech., 1975, 67, 625. 1485 SPECIAL SECTION: SOLAR PHYSICS 81. Ivanova, T. S. and Ruzmaikin, A. A., Sov. Astron., 1977, 21, 479. 82. Yoshimura, H., Astrophys. J., 1978, 226, 706. 83. Krause, F. and Meinel, R., Geophys. Astrophys. Fluid Dyn., 1988, 43, 95. 84. Weiss, N. O., Cattaneo, F. and Jones, C. A., Geophys. Astrophys. Fluid Dyn., 1984, 30, 305. 85. Jennings, R. L. and Weiss, N. O., Mon. Not. R. Astron. Soc., 1991, 252, 249. 86. Choudhuri, A. R., Astron. Astrophys., 1992, 253, 277. 87. Hoyng, P., Astron. Astrophys., 1993, 272, 321. 1486 88. Moss, D., Brandenburg, A., Tavakol, R. and Tuominen, I., Astron. Astrophys., 1992, 265, 843. 89. Ossendrijver, A. J. H., Hoyng, P. and Schmitt, D., Astron. Astrophys., 1996, 313, 938. 90. Gilman, P. A., Astrophys. J. Suppl., 1983, 53, 243. 91. Glatzmaier, G. A., Astrophys. J., 1985, 291, 300. 92. Brandenburg, A., Nordlund, A., Pulkkinen, P., Stein, R. F. and Tuominen, I., Astron. Astrophys., 1990, 232, 277. 93. Nordlund, A., Brandenburg, A., Jennings, R. L., Rieutord, M., Ruokolainen, J., Stein, R. F. and Tuominen, I., Astrophys. J., 1992, 392, 647. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Solar neutrinos: An overview J. N. Bahcall Building E, Institute for Advanced Study, Olden Lane, Princeton, NJ 08540, USA I summarize the current state of solar-neutrino research. 1. Introduction THE most important result from solar-neutrino research is, in my view, that solar neutrinos have been detected experimentally with fluxes and energies that are qualitatively consistent with solar models that are constructed assuming that the Sun shines by nuclear fusion reactions. The first experimental result1,2 has now been confirmed by four other beautiful experiments3–6. The observation of solar neutrinos with approximately the predicted energies and fluxes establishes empirically the theory7 that main-sequence stars derive their energy from nuclear fusion reactions in their interiors, and has inaugurated what we all hope will be a flourishing field of observational-neutrino astronomy. The detection of solar neutrinos settle experimentally the debate over the age and energy source of the Sun that raged for many decades, beginning in the middle of the 19th century. The leading theoretical physicists of the 19th century argued convincingly that the Sun could not be more than 107 years old because that was the maximum lifetime that could be fueled by gravitational energy (‘no other natural explanation, except chemical action, can be conceived’8). On the other hand, geologists and evolutionary biologists argued that the Sun must be > 109 years old in order to account for the observed geological features and for the evolutionary processes9. (The arguments of Lord Kelvin and his theoretical physics associates were so persuasive that in later editions Darwin removed all mention of time scales from The Origin of the Species.) Today we know that the biologists and geologists were right and the theoretical physicists were wrong, which may be a historical lesson to which we physicists should pay attention. I will discuss predictions of the combined standard model in the main part of this review. By ‘combined’ standard model I mean the predictions of the standard solar model and the predictions of the minimal electroweak theory. We need a solar model to tell us how many neutrinos of what energy are produced in the Sun and we need electroweak theory to tell us how the number and the flavour content of the neutrinos are changed as they make their way from the centre of the Sun to detectors on earth. For all practical purposes, standard electroweak theory states that nothing happens to solar neutrinos after they are created in the deep interior of the Sun. Using standard electroweak theory and fluxes from the standard solar model, one can calculate the rates of neutrino interactions in different terrestrial detectors with a variety of energy sensitivities. The combined standard model also predicts that the energy spectrum from a given neutrino source should be the same for neutrinos produced in terrestrial laboratories and in the Sun and that there should not be measurable time-dependences (other than the seasonal dependence caused by the earth’s orbit around the Sun). The spectral and temporal departures from standard model expectations are expected to be small in all currently operating experiments10 and have not yet yielded definitive results. Therefore, I will concentrate here on inferences that can be drawn by comparing the total rates observed in solarneutrino experiments with the combined standard model predictions. I will begin by reviewing in Section 2 the quantitative predictions of the combined standard solar model and then describe in Section 3 the three solar-neutrino problems that are established by the chlorine, Kamiokande, SAGE, GALLEX and Superkamiokande experiments. In Section 4 I detail the uncertainties in the standard model predictions and then show in Section 5 that helioseismological measurements indicate that the standard solar model predictions are accurate for our purposes. In Section 5 I discuss the implications for solar-neutrino research of the precise agreement between helioseismological measurements and the predictions of standard solar models. Next, ignoring all knowledge of the Sun, I cite analyses in Section 6 that show that one cannot fit the existing experimental data with neutrino fluxes that are arbitrary parameters, unless one invokes new physics to change the shape or flavour content of neutrino energy spectrum. I summarize in Section 7 the characteristics of the best-fitting neutrino oscillation descriptions of the experimental data. Finally, I will discuss and summarize results in Section 8. If you want to obtain numerical data or subroutines that are discussed in this review, or to see relevant background information, you can copy them from my Web site: http://www.sns.ias.edu/~ jnb. 2. Standard model predictions Table 1 gives the neutrino fluxes and their uncertainties for our best standard solar model11, hereafter BP98. Figure 1 shows the predicted neutrino fluxes from the dominant p–p fusion chain. e-mail: jnb@ias.edu CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1487 SPECIAL SECTION: SOLAR PHYSICS The BP98-solar model includes diffusion of heavy elements and helium, makes use of the nuclear reaction rates recommended by the expert workshop held at the Institute of Nuclear Theory12, recent (1996) Livermore OPAL radiative opacities13, the OPAL equation of state14, and electron and ion screening as determined Table 1. Standard model predictions (BP98): solar-neutrino fluxes and neutrino capture rates, with 1σuncertainties from all sources (combined quadratically) Source p–p Flux (10 10 cm–2 s–1 ) + 0. 01 5.94 (1.00 − 0.01 ) Cl (SNU) Ga (SNU) 0.0 69.6 pep 1.39 × 10 (1 .00 − 0.01 ) 0.2 2.8 hep 2.10 × 10 –7 0.0 0.0 7 Be 4.80 × 10 –1 (1 .00 − 0.09 ) 1.15 34.4 8 B 5.15 × 10 –4 (1 .00 − 0.14 ) 5.9 12.4 –2 + 0. 01 + 0. 09 + 0. 19 N 6.05 × 10 (1 .00 − 0.13 ) 0.1 3.7 15 O 5.32 × 10 –2 (1 .00 − 0.15 ) 0.4 6.0 F + 0. 12 10 –4 (1 .00 − 0.11 ) 0.0 0.1 17 Total –2 + 0. 19 13 + 0. 22 6.33 × + 1. 2 7.7 − 1.0 129 +8 −6 SNU is a unit used to describe the measured rates of solar-neutrino radiochemical experiments (10–36 interactions per target atom per second). 1488 by the recent density matrix calculation15,16. The neutrino absorption cross-sections that are used in constructing Table 1 are the most accurate values available17,18 and include, where appropriate, the thermal energy of fusing-solar ions and improved nuclear and atomic data. The validity of the absorption cross-sections has recently been confirmed experimentally using intense radioactive sources of 51Cr. The ratio, R, of the capture rate measured (in GALLEX and SAGE) to the calculated 51Cr-capture rate is R = 0.95 ± 0.07 (exp)+ (theory) and was discussed extensively at Neutrino 98 by Gavrin and by Kirsten. The neutrino–electron scattering cross-sections, used in interpreting the Kamiokande and + 0. 04 SuperKamiokande experiments, now include electroweak radiative − 0. 03 corrections19. Figure 2 shows for the chlorine experiment all the predicted rates and the estimated uncertainties (1σ) published by my colleagues and myself since the first measurement by Ray Davis and his colleagues in 1968. This figure should give you some feeling for the robustness of the solar model calculations. Many hundreds and probably thousands of researchers have, over three decades, made great improvements in the input data for the solar models, including nuclear cross-sections, neutrino cross-sections, measured element abundances on the surface of the Sun, the solar luminosity, the stellar radiative opacity, and the stellar equation of state. Nevertheless, the most accurate predictions of today are Figure 1. The energy spectrum of neutrinos from the p–p chain of interactions in the Sun, as predicted by the standard solar model. Neutrino fluxes from continuum sources (such as p–p and 8 B) are given in the units of counts per cm2 per second. The p–p chain is responsible for more than 98% of the energy generation in the standard solar model. Neutrinos produced in the carbon– nitrogen–oxygen (CNO) chain are not important energetically and are difficult to detect experimentally. The arrows at the top of the figure indicate the energy thresholds for the ongoing neutrino experiments. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS essentially the same as they were in 1968 (although now they can be made with much greater confidence). For the gallium experiments, the neutrino fluxes predicted by standard solar models, corrected for diffusion, have been in the range 120 SNU to 141 SNU since 1968 (ref. 17): A SNU is a convenient unit with which to describe the measured rates of solar-neutrino experiments; 10–36 interactions per target atom per second. There are three reasons that the theoretical calculations of neutrino fluxes are robust: (i) the availability of precision measurements and precision calculations of input data; (ii) the connection between neutrino fluxes and the measured solar luminosity; and (iii) the measurement of the helioseismological frequencies of the solar pressure- mode (p-mode) eigenfrequencies. I have discussed these reasons in detail elsewhere20. Figure 3 shows the calculated 7Be- and 8B-neutrino fluxes for all 19 standard solar models which have been published in the last 10 years in refereed science journals. The fluxes are normalized by dividing each published value by the flux from the BP98-solar model11: the abscissa is the normalized-8B flux and the ordinate is the normalized-7Be neutrino flux. The rectangular box shows the estimated 3σ uncertainties in the predictions of the BP98 solar model. All of the solar model results from different groups fall within the estimated 3σ uncertainties in the BP98 analysis (with one not understood exception that falls slightly outside). This agreement demonstrates the robustness of the predictions, since the calculations use different computer codes (which achieve varying degrees of precision) and involve a variety of choices for the nuclear parameters, the equation of state, the stellar radiative opacity, the initial heavy element abundances, and the physical processes that are included. The largest contributions to the dispersion in values in Figure 3 Figure 2. The predictions of John Bahcall and his collaborators of neutrino-capture rates in the 37 Cl experiment are shown as a function of the date of publication (since the first experimental report1 in 1968). The event rate, SNU, is a convenient product of neutrinoflux times the interaction cross-section, 10–36 interactions per target atom per sec. The format is from Figure 1.2 (ref. 40). The predictions have been updated through 1998. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 are due to the choice of the normalization for S 17 (the production cross-section factor for 8B neutrinos) and the inclusion or noninclusion of element diffusion in the stellar-evolution codes. The effect in the plane of Figure 3 of the normalization of S 17 is shown by the difference between the point for BP98 (1.0, 1.0), which was computed using the most recent recommended normalization12, and the point at (1.18, 1.0) which corresponds to the BP98 result with the earlier (CalTech) normalization21. Helioseismological observations have shown11,22 that element diffusion is occurring and must be included in solar models, so that the most recent models shown in Figure 3 now all include helium and heavy-element diffusion. By comparing a large number of earlier models, it was shown that all published standard solar models give the same results for solar-neutrino fluxes to an accuracy of better than 10% if the same input parameters and physical processes are included23,24. Bahcall et al.10 have compared the observed rates with the calculated standard-model values, combining quadratically the theoretical solar model and experimental uncertainties, as well as the uncertainties in the neutrino cross-sections. Since the GALLEX and SAGE experiments measure the same quantity, we treat the weighted average rate in gallium as one experimental number. We adopt the SuperKamiokande measurement as the most precise direct determination of the higher-energy 8 B-neutrino flux. Using the predicted fluxes from the BP98 model, the χ2 for the Figure 3. Predictions of standard solar models since 1988. This figure, which is Figure 1 of ref. 10, shows the predictions of 19 standard solar models in the plane defined by the 7 Be in and 8 B neutrino fluxes. The abbreviations that are used in the figure to identify different solar models are defined in the bibliographical item, ref. 45. The figure includes all standard solar models, with which I am familiar that were published in refereed journals in the decade 1988– 1998. All of Figure the fluxes 4. are Comparison normalizedofto measured the predictions rates of andthestandard-model 2–6 Bahcall–Pinsonneault predictions 1998 for solar five solar-neutrino model, BP9811experiments . The rectangular . The unit for the error box defines radiochemical the 3σ-errorexperiments range of the (chlorine BP98 fluxes. and gallium) The best-fit is SNU (see Figure 7 Be-neutrino flux 2 for is negative. a definition); At thethe 99% unit C.L., forthere the is water-Cerenkov no solution10 experiments with all positive (Kamiokande neutrino fluxes and (see SuperKamiokande) discussion in section is the6).rate Allpredicted of by the the standard model standard solutions solar model lie farplus from thethe standard best-fitelectroweak solution, even theory11 . far from the 3σ contour. 1489 SPECIAL SECTION: SOLAR PHYSICS fit to the three experimental rates (chlorine, gallium, and SuperKamiokande, see Figure 4) is: 2 χSSM (3 experimental rates) = 61. (1) The result given in eq. (1), which is approximately equivalent to a 20σ discrepancy, is a quantitative expression of the fact that the standard model predictions do not fit the observed solar-neutrino measurements. 3. Three solar-neutrino problems I will now compare the predictions of the combined standard model with the results of the operating solar-neutrino experiments. We will see that this comparison leads to three different discrepancies between the calculations and the observations, which I will refer to as the three solar-neutrino problems. Figure 4 shows the measured and the calculated event rates in the five ongoing solar-neutrino experiments. This figure reveals three discrepancies between the experimental results and the expectations based upon the combined standard model: As we shall see, only the first of these discrepancies depends in significant measure upon the predictions of the standard solar model. Calculated vs observed absolute rate The first solar-neutrino experiment to be performed was the chlorine radiochemical experiment2, which detects electron-type neutrinos that are more energetic than 0.81 MeV. After more than a quarter of a century of operation of this experiment, the measured event rate is 2.56 ± 0.23 SNU, which is a factor of three less than predicted by most detailed theoretical calculations, 7.7 +−11..02 SNU (ref. 11). Most of the predicted rate in the chlorine experiment is from the rare, high-energy 8B neutrinos, although the 7 Be neutrinos are also expected to contribute significantly. According to standard-model calculations, the pep neutrinos and the CNO neutrinos (for simplicity not discussed here) are expected to contribute less than 1 SNU to the total event rate. This discrepancy between the calculations and the observations for the chlorine experiment was for more than two decades, the only solar-neutrino problem. I shall refer to the chlorine disagreement as the first solar-neutrino problem. Incompatibility of chlorine and water experiments The second solar-neutrino problem results from a comparison of the measured event rates in the chlorine experiment and in the Japanese pure-water experiments, Kamiokande3 and 6 SuperKamiokande . The water experiments detect higher-energy neutrinos, most easily above 7 MeV, by observing the Cerenkov radiation from neutrino–electron scattering: ν + e → ν′ + e′. According to the standard solar model, 8B-beta decay, and 1490 possibly the hep reaction25, are the only important source of these higher-energy neutrinos. The Kamiokande and SuperKamiokande experiments show that the observed neutrinos come from the Sun. The electrons that are scattered by the incoming neutrinos recoil predominantly in the direction of the Sun–Earth vector; the relativistic electrons are observed by the Cerenkov radiation they produce in the water detector. In addition, the water Cerenkov experiments measure the energies of individual scattered electrons and therefore provide information about the energy spectrum of the incident solar neutrinos. The total event rate in the water experiments, about 0.5 the standard-model value (see Figure 4), is determined by the same high-energy 8B neutrinos that are expected, on the basis of a combined standard model, to dominate the event rate in the chlorine experiment. I have shown elsewhere26 that solar physics changes the shape of the 8B-neutrino spectrum by less than 1 part in 105. Therefore, we can calculate the rate in the chlorine experiment (threshold 0.8 MeV) that is produced by the 8B neutrinos observed in the Kamiokande and SuperKamiokande experiments at an order of magnitude higher energy threshold. If no new physics changes the shape of the 8B-neutrino energy spectrum, the chlorine rate from 8B alone is 2.8 ± 0.1 SNU for the SuperKamiokande normalization (3.2 ± 0.4 SNU for the Kamiokande normalization), which exceeds the total observed chlorine rate of 2.56 ± 0.23 SNU. Comparing the rates of the SuperKamiokande and the chlorine experiments, one finds – assuming that the shape of the energy spectrum of 8Bνe ’s is not changed by new physics – that the net contribution to the chlorine experiment from the pep, 7Be and CNO neutrino sources is negative: – 0.2 ± 0.3 SNU. The contributions from the pep, 7 Be, and CNO neutrinos would appear to be completely missing; the standard model prediction for the combined contribution of pep, 7Be, and CNO neutrinos is a relatively large 1.8 SNU (see Table 1). On the other hand, we know that the 7Be neutrinos must be created in the Sun since they are produced by electron capture on the same isotope ( 7Be) which gives rise to the 8B neutrinos by proton capture. Hans Bethe and I pointed out27 that this apparent incompatibility of the chlorine and water-Cerenkov experiments constitutes a second solar-neutrino problem that is almost independent of the absolute rates predicted by solar models. The inference that is usually made from this comparison is that the energy spectrum of 8B neutrinos is changed from the standard shape by physics not included in the simplest version of the standard electroweak model. Gallium experiments: No room for 7Be neutrinos The results of the gallium experiments, GALLEX and SAGE, constitute the third solar-neutrino problem. The average observed rate in these two experiments is 73 ± 5 SNU, which is accounted for in the standard model by the theoretical rate of 72.4 SNU that CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS the calculated 7Be-solar-neutrino flux. The required change in the nuclear-physics cross-section would also increase the predicted neutrino event rate by more than 100 in the Kamiokande experiment, making that prediction completely inconsistent with what is observed. I conclude that either: (i) at least three of the five operating solar-neutrino experiments (the two gallium experiments plus either chlorine or the two water-Cerenkov experiments, Kamiokande and SuperKamiokande) have yielded misleading results, or (ii) physics beyond the standard electroweak model is required to change the energy spectrum of νe after the neutrinos are produced in the centre of the Sun. is calculated to come from the basic p–p and pep neutrinos (with only a 1% uncertainty in the standard solar model p–p flux). The 8B neutrinos, which are observed above 6.5 MeV in the Kamiokande experiment, must also contribute to the gallium event rate. Using the standard shape for the spectrum of 8B neutrinos and normalizing to the rate observed in Kamiokande, 8B contributes another 6 SNU. (The contribution predicted by the standard model is 12 SNU, see Table 1.) Given the measured rates in the gallium experiments, there is no room for the additional 34 ± 3 SNU that is expected from 7Be neutrinos on the basis of standard solar models (see Table 1). The seeming exclusion of everything but p–p neutrinos in the gallium experiments is the third solar-neutrino problem. This problem is essentially independent of the previously-discussed solar-neutrino problems, since it depends strongly upon the p–p neutrinos that are not observed in the other experiments and whose theoretical flux can be calculated accurately. The missing 7Be neutrinos cannot be explained away by a change in solar physics. The 8B neutrinos that are observed in the Kamiokande experiment are produced in competition with the missing 7Be neutrinos; the competition is between electron capture on 7Be vs proton capture on 7Be. Solar model explanations that reduce the predicted 7Be flux generically reduce much more (too much) the predictions for the observed 8B flux. The flux of 7Be neutrinos, φ( 7Be), is independent of measurement uncertainties in the cross-section for the nuclear reaction 7Be(p, γ)8B; the cross-section for this proton-capture reaction is the most uncertain quantity that enters in an important way in the solar model calculations. The flux of 7Be neutrinos depends upon the proton-capture reaction only through the ratio φ ( 7 Be) ∝ R (e) , R( e ) + R( p ) 4. Uncertainties in the flux calculations I will now discuss uncertainties in the solar-model-flux calculations. Table 2 summarizes the uncertainties in the most important solar-neutrino fluxes and in the Cl and Ga event rates due to different nuclear fusion reactions (the first four entries), the heavy element to hydrogen mass ratio (Z/X), the radiative opacity, the solar luminosity, the assumed solar age, and the helium and heavy element diffusion coefficients. The 14N + p reaction causes a 0.2% uncertainty in the predicted p–p flux and a 0.1 SNU uncertainty in the Cl (Ga) event rates. The predicted event rates for the chlorine and gallium experiments use recent improved calculations of neutrinoabsorption cross-sections17,18. The uncertainty in the prediction for the gallium rate is dominated by uncertainties in the neutrinoabsorption cross sections, + 6.7 SNU (7% of the predicted rate) and – 3.8 SNU (3% of the predicted rate). The uncertainties in the chlorine-absorption cross- sections cause an error, ± 0.2 SNU (3% of the predicted rate), that is relatively small compared to other uncertainties in predicting the rate for this experiment. For nonstandard neutrino-energy spectra that result from new neutrino physics, the uncertainties in the predictions for currently favoured solutions (which reduce the contributions from the least welldetermined 8B neutrinos) will in general be less than the values quoted here for standard spectra and must be calculated using the appropriate cross-section uncertainty for each neutrino (2) where R(e) is the rate of electron capture by 7Be nuclei and R(p) is the rate of proton capture by 7Be. With standard parameters, solar models yield R(p) ≈ 10–3R(e). Therefore, one would have to increase the value of the 7Be(p, γ)8B cross-section by more than two orders of magnitude over the current-best estimate (which has an estimated uncertainty of ~ 10%) in order to affect significantly Table 2. Average uncertainties in neutrino fluxes and event rates due to different input data. The flux uncertainties are expressed in fractions of the total flux, and the event-rate uncertainties are expressed in SNU. The 7 Be-electron capture rate causes an uncertainty of ± 2% (ref. 44) that affects only the 7 Be-neutrino flux. The average fractional uncertainties for individual parameters are shown Fractional uncertainty p–p 0.017 3 He3 He 0.060 3 He4 He 0.094 7 Be + p 0.106 Z/X Opacity Luminocity 0.004 Age 0.004 0.003 0.003 0.0 0.003 0.028 0.014 0.003 0.018 0.052 0.028 0.006 0.040 Diffuse 0.033 Flux p–p 0.002 0.002 0.005 0.000 0.002 7 8 Be B 0.023 0.0155 0.040 0.080 0.000 0.019 0.021 0.075 0.105 0.042 CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1491 SPECIAL SECTION: SOLAR PHYSICS energy17,18. The nuclear fusion uncertainties in Table 2 were taken from Adelberger et al.12, the neutrino cross-section uncertainties from refs 17, 18, the heavy element uncertainty was taken from helioseismological measurements28, the luminosity and age uncertainties were adopted from BP95 (ref. 24), the 1σ-fractional uncertainty in the diffusion rate was taken to be 15% (ref. 29), which is supported by helioseismological evidence22, and the opacity uncertainty was determined by comparing the results of fluxes computed using the older Los Alamos opacities with fluxes computed using the modern Livermore opacities23. To include the effects of asymmetric errors, the now publicly-available code for calculating rates and uncertainties (see discussion in previous section) was run with different input uncertainties and the results averaged. The software contains a description of how each of the uncertainties listed in Table 2 were determined and used. The low-energy cross-section of the 7Be + p reaction is the most important quantity that must be determined more accurately in order to decrease the error in the predicted event rates in solarneutrino experiments. The 8B-neutrino flux that is measured by the Kamiokande3, Super-Kamiokande6, and SNO30 experiments is, in all standard solar model calculations, directly proportional to the 7Be + p cross-section. If the 1σ uncertainty in this cross-section can be reduced by a factor of two to 5%, then it will no longer be the limiting uncertainty in predicting the crucial 8B-neutrino flux (cf. Table 2). 5. How large an uncertainty does helioseismology suggest? Could the solar model calculations be wrong by enough to explain the discrepancies between predictions and measurements for solarneutrino experiments? Helioseismology, which confirms predictions of the standard solar model to high precision, suggests that the answer is probably ‘No’. Figure 5 shows the fractional differences between the most accurate available sound speed measured by helioseismology31 and sound speed calculated with our best solar model (with no free parameters). The horizontal line corresponds to the hypothetical case in which the model predictions exactly match the observed values. The root mean square (rms) fractional difference between the calculated and the measured sound speeds is 1.1 × 10–3 for the entire region over which the sound speeds are measured, 1492 0.05R¤ < R < 0.95R¤. In the solar core, 0.05R¤ < R < 0.25R¤ (in which about 95% of the solar energy and neutrino flux is produced in a standard model), the rms fractional difference between measured and calculated sound speeds is 0.7 × 10–3. Helioseismological measurements also determine two other parameters that help characterize the outer part of the Sun (far from the inner region in which neutrinos are produced): the depth of the solar convective zone (CZ), the region in the outer part of the Sun that is fully convective and the present-day surfaceabundance by mass of helium (Ysurf). The measured values, RCZ = (0.713 ± 0.001)R¤ (ref. 32), and Ysurf = 0.249 ± 0.003 (ref. 28), are in satisfactory agreement with the values predicted by the solar model BP98, namely, RCZ = 0.714R¤, and Ysurf = 0.243. However, we shall see below that precision measurements of the sound speed near the transition between the radiative interior (in which energy is transported by radiation) and the outer convective zone (in which energy is transported by convection) reveal small discrepancies between the model predictions and the observations in this region. If solar physics were responsible for the solar-neutrino problems, how large would one expect the discrepancies to be between the solar model predictions and helioseismological observations? The characteristic size of the discrepancies can be estimated using the results of the neutrino experiments and scaling laws for neutrino fluxes and sound speeds. All recently published solar models predict essentially the same fluxes from the fundamental p–p and pep reactions (amounting to 72.4 SNU in gallium experiments, cf. Table 1), which are closely related to the solar luminosity. Comparing the measured gallium rates and the standard predicted rate for the gallium experiments, the 7Be flux must be reduced by a factor N if the disagreement is not to exceed n standard deviations, where N and n satisfy 72.4 + (34.4)/N = 72.2 + nσ. For a 1σ (3σ) disagreement, N = 6.1 (2.05). Sound-speeds scale like the square root of the local temperature divided by the mean molecular weight and the 7Beneutrino flux scales approximately as the 10th power of the temperature33. Assuming that the temperature changes are dominant, agreement to within 1σ would require fractional changes of order 0.09 in sound speeds (3σ could be reached with 0.04 changes), if all model changes were in the temperature.* This argument is conservative because it ignores the 8B and CNO neutrinos which contribute to the observed counting rate (cf. Table 1) and which, if included, would require an even larger reduction of the 7Be flux. *I have used in this calculation the GALLEX and SAGE measured rates reported by Kirsten and Gavrin at Neutrino 98. The experimental rates used in BP98 were not as precise and therefore resulted in slightly less stringent constraints than those imposed here. In BP98, we found that agreement to within 1σ with the then available experimental numbers would require fractional changes of order 0.08 in sound speeds (3σcould be reached with 0.03 changes.) CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS I have chosen the vertical scale in Figure 5 to be appropriate for fractional differences between measured and predicted sound speeds that are of order 0.04 to 0.09, and that might therefore affect solar-neutrino calculations. Figure 5 shows that the characteristic agreement between the solar model predictions and helioseismological measurements is more than a factor of 40 better than would be expected if there were a solar model explanation of the solar-neutrino problems. 6. Fits without solar models Suppose (following the precepts of Hata et al.34, Parke35, and Heeger and Robertson36) we now ignore everything we have learned about solar models over the last 35 years and allow the important p–p, 7Be, and 8B fluxes to take on any non-negative values. What is the best fit that one can obtain to the solarneutrino measurements assuming only that the luminosity of the Sun is supplied by nuclear fusion reactions among light elements (the so-called ‘luminosity constraint’37)? The answer is that the fits are bad, even if we completely ignore what we know about the Sun; I quote the results from ref. 10. If the CNO-neutrino fluxes are set equal to zero, there are no acceptable solutions at the 99% C. L. (~ 3σ result). The best-fit is worse if the CNO fluxes are not set equal to zero. All so-called ‘solutions’ of the solar-neutrino problems in which the astrophysical model is changed arbitrarily (ignoring helioseismology and other constraints) are inconsistent with the observations at much more than a 3σ level of significance. No fiddling of the physical conditions in the model can yield the minimum value, quoted above, that was found by varying the fluxes independently and arbitrarily. Figure 3 shows, in the lower left-hand corner, the best-fit solution and the 1σ – 3σ contours. The 1σ and 3σ limits were obtained by requiring that χ2 = χ2min + δχ2, where for 1σ, δχ2 = 1 and for 3σ, δχ2 = 9. All of the standard model solutions lie far from the best-fit solution and even lie far from the 3σ contour. Since standard model descriptions do not fit the solar-neutrino data, we will now consider models in which neutrino oscillations change the shape of the neutrino energy spectra. Table 3. Neutrino oscillation solutions Solution SMA LMA LOW VAC 5 2 8 8 ∆m 2 (eV2 ) sin2 2θ × 10 –6 × 10 –5 × 10 –8 × 10 –11 5 × 10 –3 0.8 0.96 0.7 production-reaction cross-section (3He + p → 4He + e+ + νe ) is used10. However, for over a decade I have not given an estimated uncertainty for this cross-section40. The transition matrix element is essentially forbidden and the actual quoted value for the production cross-section depends upon a delicate cancellation between two comparably sized terms that arise from very different and hard to evaluate nuclear physics. I do not see any way at present to determine from experiment or from first principles theoretical calculations a relevant, robust upper limit to the hep-production cross-section (and therefore the hep solar-neutrino flux). The possible role of hep neutrinos in solar-neutrino experiments is discussed extensively in ref. 25. The most important unsolved problem in theoretical nuclear physics related to solar neutrinos is the range of values allowed by fundamental physics for the hep-production cross-section. 8. Discussion and conclusion When the chlorine solar-neutrino experiment was first proposed41, the only stated motivation was ‘. . . to see into the interior of a star and thus verify directly the hypothesis of nuclear energy generation in stars’. This goal has now been achieved. The focus has shifted to using solar-neutrino experiments as a tool for learning more about the fundamental characteristics of 7. Neutrino oscillations The experimental results from all five of the operating solarneutrino experiments (chlorine, Kamiokande, SAGE, GALLEX, and SuperKamiokande) can be fit well by descriptions involving neutrino oscillations, either vacuum oscillations (as originally suggested by Gribov and Pontecorvo38) or resonant matter oscillations (as originally discussed by Mikeyhev, Smirnov, and Wolfenstein (MSW)39). Table 3 summarizes the four best-fit solutions that are found in the two-neutrino approximation10,25. Only the SMA MSW solution fits well all the data – including the recoil electron energy spectrum measured in the SuperKamiokande experiment – if the standard value for the hep CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 Figure 5. Predicted vs measured sound speeds. This figure shows the excellent agreement between the calculated (solar model BP98, Model) and the measured (Sun) sound speeds, a fractional difference of 0.001 rms for all speeds measured between 0.05R¤ and 0.95R¤. The vertical scale is chosen so as to emphasize that the fractional error is much smaller than generic changes in the model, 0.04 to 0.09, that might significantly affect the solar-neutrino predictions. 1493 SPECIAL SECTION: SOLAR PHYSICS neutrinos as particles. Experimental effort is now concentrated on answering the question: What are the probabilities for transforming a solar νe of a definite energy into the other possible neutrino states? Once this question is answered, we can calculate what happens to νe ’s that are created in the interior of the Sun. Armed with this information from weak interaction physics, we can return again to the original motivation of using neutrinos to make detailed, quantitative tests of nuclear fusion rates in the solar interior. Measurements of the flavour content of the dominant low energy neutrino sources, p–p and 7Be neutrinos, will be crucial in this endeavour and will require another generation of superb solarneutrino experiments. Three decades of refining the input data and the solar model calculations has led to a predicted standard model event rate for the chlorine experiment, 7.7 SNU, which is very close to 7.5 SNU, the best-estimate value obtained in 1968 (ref. 42). The situation regarding solar neutrinos is, however, completely different now, thirty years later. Four experiments have confirmed the original chlorine detection of solar neutrinos. Helioseismological measurements are in excellent agreement with the standard solar model predictions and very strongly disfavour (by a factor of 40 or more) hypothetical deviations from the standard model that are required to fit the neutrino data (cf. Figure 5). Just in the last two years, improvements in the helioseismological measurements have resulted in a five-fold improvement in the agreement between the calculated standard solar model sound speeds and the measured solar velocities. 1. Davis, R. Jr., Harmer, D. S. and Hoffman, K. C., Phys. Rev. Lett., 1968, 20, 1205–1209. 2. Davis, R. Jr., Prog. Part. Nucl. Phys., 1994, 32, 13; Cleveland, B. T., Daily, T., Davis, R. Jr., Distel, J. R., Lande, K., Lee, C. K., Wildenhain, P. S. and Ullman, J., Astrophys. J., 1998, 496, 505–526. 3. Kamiokande Collaboration, Fukuda, Y. et al., Phys. Rev. Lett., 1996, 77, 1683–1686. 4. SAGE Collaboration, Gavrin, V. et al., in Proceedings of the 17th International Conference on Neutrino Physics and Astrophysics (Helsinki) (eds Enqvist, K., Huitu, K. and Maalampi, J.), World Scientific, Singapore, 1997, pp. 14–24; SAGE collaboration, Abdurashitov, J. N. et al., Phys. Rev. Lett., 1996, 77, 4708–4711. 5. GALLEX Collaboration, Anselmann, P. et al., Phys. Lett., 1995, B342, 440–450; GALLEX Collaboration, Hampel, W. et al., Phys. Lett., 1996, B388, 384–396. 6. SuperKamiokande Collaboration, Suzuki, Y., in Proceedings of the XVIII International Conference on Neutrino Physics and Astrophysics, Takayama, Japan, 4–9 June 1998 (eds Suzuki, Y. and Totsuka, Y.), to be published in Nucl. Phys. (Proc. Suppl); the Super-Kamiokande Collaboration, Fukuda, Y. et al., Phys. Rev. Lett., 1998, B81, 1562–1567; Fukuda, Y. et al., hepex/9812009; Totsuka, Y., in the Proceedings of the 18th Texas Symposium on Relativistic Astrophysics and Cosmology, 15–20 December 1996, Chicago, Illinois (eds Olinto, A., Frieman, J. and Schramm, D.), World Scientific, Singapore, 1998, pp. 114– 123. 7. Bethe, H. A., Phys. Rev., 1939, 55, 434–456. 8. Thompson, W., On the Age of the Sun’s Heat, MacMilan’s Magazine, 1862, 5, 288–393; reprinted in Treatise on Natural Philosphy (eds Thompson, W. and Gutherie, P.), Cambridge Univ. Press, Cambridge, 1883, vol. 1, appendix E, pp. 485–494. 1494 9. Darwin, C., On the Origin of the Species, 1859 (Reprinted: Harvard University Press, Cambridge, MA, 1964). 10. Bahcall, J. N., Krastev, P. I. and Smirnov, A. Yu., Phys. Rev., 1998, D58, 096016–096022. 11. Bahcall, J. N., Basu, S. and Pinsonneault, M. H., Phys. Lett., 1998, B433, 1–8. 12. Adelberger, E. et al., Rev. Mod. Phys., 1998, 70, 1265–1292. 13. Iglesias, C. A. and Rogers, F. J., Astrophys. J., 1996, 464, 943– 953; Alexander, D. R. and Ferguson, J. W., Astrophys. J., 1994, 437, 879–891. These references describe the different versions of the OPAL opacities. 14. Rogers, F. J., Swenson, F. J. and Iglesias, C. A., Astrophys. J., 1996, 456, 902–908. 15. Gruzinov, A. V. and Bahcall, J. N., Astrophys. J., 1998, 504, 996–1001. 16. Salpeter, E. E., Austr. J. Phys., 1954, 7, 373–388. 17. Bahcall, J. N., Phys. Rev., 1997, C56, 3391–3409. 18. Bahcall, J. N., Lisi, E., Alburger, D. E., De Braeckeleer, L., Freedman, S. J. and Napolitano, J., Phys. Rev., 1996, C54, 411– 422. 19. Bahcall, J. N., Kamionkowski, M. and Sirlin, A., Phys. Rev., 1995, D51, 6146–6158. 20. Bahcall, J. N., Astrophys. J., 1996, 467, 475–484. 21. Johnson, C. W., Kolbe, E., Koonin, S. E. and Langanke, K., Astrophys. J., 1992, 392, 320–327. 22. Bahcall, J. N., Pinsonneault, M. H., Basu, S. and ChristensenDalsgaard, J., Phys. Rev. Lett., 1997, 78, 171–174. 23. Bahcall, J. N. and Pinsonneault, M. H., Rev. Mod. Phys., 1992, 64, 885–926. 24. Bahcall, J. N. and Pinsonneault, M. H., Rev. Mod. Phys., 1995, 67, 781–808. 25. Bahcall, J. N. and Krastev, P. I., Phys. Lett., 1998, B436, 243– 250. 26. Bahcall, J. N., Phys. Rev., 1991, D44, 1644–1651. 27. Bahcall, J. N. and Bethe, H. A., Phys. Rev. Lett., 1990, 65, 2233–2235. 28. Basu, S. and Antia, H. M., Mon. Not. R. Astron. Soc., 1997, 287, 189–198. 29. Thoul, A. A., Bahcall, J. N. and Loeb, A., Astrophys. J., 1994, 421, 828–842. 30. McDonald, A. B., in The Proceedings of the 9th Lake Louise Winter Institute (eds Astbury, A. et al.), World Scientific, Singapore, 1994, pp. 1–22. 31. Basu, S. et al., Mon. Not. R. Astron. Soc., 1997, 292, 234–251. 32. Basu, S. and Antia, H. M., Mon. Not. R. Astron. Soc., 1995, 276, 1402–1408. 33. Bahcall, J. N. and Ulmer, A., Phys. Rev., 1996, D53, 4202– 4210. 34. Hata, N., Bludman, S. and Langacker, P., Phys. Rev., 1994, D49, 3622–3625. 35. Parke, S., Phys. Rev. Lett., 1995, 74, 839–841. 36. Heeger, K. M. and Robertson, R. G. H., Phys. Rev. Lett., 1996, 77, 3720–3723. 37. Bahcall, J. N. and Krastev, P. I., Phys. Rev., 1996, D53, 4211– 4225. 38. Gribov, V. N. and Pontecorvo, B. M., Phys. Lett., 1969, B28, 493–496; Pontecorvo, B., Sov. Phys. JETP, 1968, 26, 984– 988. 39. Wolfenstein, L., Phys. Rev., 1978, D17, 2369–2374; Mikheyev, S. P. and Smirnov, A. Yu., Yad. Fiz., 1985, 42, 1441–1448 (Sov. J. Nucl. Phys., 1985, 42, 913–917); Nuovo Cimento, 1986, C9, 17–26. 40. Bahcall, J. N., Neutrino Astrophysics, Cambridge University Press, Cambridge, 1989. 41. Bahcall, J. N., Phys. Rev. Lett., 1964, 12, 300–302; Davis, R. Jr., Phys. Rev. Lett., 1964, 12, 303–307; Bahcall, J. N. and Davis, R. Jr., in Stellar Evolution (eds Stein, R. F. and Cameron, A. G.), Plenum Press, New York, 1966, pp. 241–243 [proposal first CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS made in 1963 at this conference]. 42. Bahcall, J. N., Bahcall, N. A. and Shaviv, G., Phys. Rev. Lett., 1968, 20, 1209–1212. 43. Bahcall, J. N. and Pinsonneault, M. H., in Proceedings of the 17th International Conference on Neutrino Physics and Astrophysics (Helsinki) (eds Enqvist, K., Huitu, K. and Maalampi, J.), World Scientific, Singapore, 1997, pp. 56–70. 44. Gruzinov, A. V. and Bahcall, J. N., Astrophys. J., 1997, 490, 437–441. 45. (GONG) Christensen-Dalsgaard, J. et al., Science, 1996, 272, 1286–1292; (BP95) Bahcall, J. N. and Pinsonneault, M. H., Rev. Mod. Phys., 1995, 67, 781–808; (KS94) Kovetz, A. and Shaviv, G., Astrophys. J., 1994, 426, 787–800; (CDF94) Castellani, V., Degl’Innocenti, S., Fiorentini, G., Lissia, L. M. and Ricci, B., Phys. Lett., 1994, B324, 425–432; (JCD94) ChristensenDalsgaard, J., Europhys. News, 1994, 25, 71–75; (SSD94) Shi, X., Schramm, D. N. and Dearborn, D. S. P., Phys. Rev., 1994, D50, 2414–2420; (DS96) Dar, A. and Shaviv, G., Astrophys. J., CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1996, 468, 933–946; (CDF93) Castellani, V., Degl’Innocenti, S. and Fiorentini, G.. Astron. Astrophys., 1993, 271, 601–620; (TCL93) Turck-Chièze, S. and Lopes, I., Astrophys. J., 1993, 408, 347–367; (BPML93) Berthomieu, G., Provost, J., Morel, P. and Lebreton, Y., Astron. Astrophys., 1993, 268, 775–791; (BP92) Bahcall, J. N. and Pinsonneault, M. H., Rev. Mod. Phys., 1992, 64, 885–926; (SBF90) Sackman, I.-J., Boothroyd, A. I. and Fowler, W. A., Astrophys. J., 1990, 360, 727–736; (BU88) Bahcall, J. N. and Ulrich, R. K., Rev. Mod. Phys., 1988, 60, 297–372; (RVCD96) Richard, O., Vauclair, S., Charbonnel, C. and Dziembowski, W. A., Astron. Astrophys., 1996, 312, 1000– 1011; (CDR97) Ciacio, F., Degl’Innocenti, S. and Ricci, B., Astron. Astrophys. Suppl. Ser., 1997, 123, 449–454. ACKNOWLEDGEMENT. #PHY95-13835. I acknowledge support from NSF grant 1495 SPECIAL SECTION: SOLAR PHYSICS The dynamics and heating of the quiet solar chromosphere Wolfgang Kalkofen*,† and Peter Ulmschneider** *Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA02138, USA **Institut für Theoretische Astrophysik, Universität Heidelberg, Tiergartenstr. 15, Germany The solar chromosphere can be characterized by two signatures: the spectroscopic signature is an emission spectrum for all radiation originating in the chromosphere; only NLTE effects in the cores of strong lines producing absorption features. And the dynamical signature is in the form of oscillations, with a period of 3 min in the nonmagnetic chromosphere and a period of 7 min in magnetic regions. The paper explains these signatures in terms of waves: The dynamics of the chromosphere is due to acoustic waves in the magnetic-field-free atmosphere, which produce K2v bright points, and to kink and longitudinal waves in magnetic flux tubes, which produce network bright points. The heating of the chromosphere is caused by acoustic waves whose dissipation makes the kinetic temperature rise in the outward direction, producing the emission spectrum. As far as energy fluxes are concerned, the energy dissipated in chromospheric heating outweighs the energy visible in bright points by two orders of magnitude. The paper interprets the observed oscillation periods in the chromosphere as cutoff periods: for the 3 min period, as the cutoff period of acoustic waves in a nonmagnetic, stratified atmosphere; and for the 7-min period, as the cutoff period of kink waves in magnetic flux tubes for field strengths typical of the magnetic network. preference for the blue peaks3. The bifurcation of the chromosphere that characterizes optical lines is also found in the ultraviolet, where SUMER shows 3-min waves in emission lines and continua of neutral metals from the internetwork chromosphere4, and 7-min waves in emission of the Lyman continuum and several Lyman lines from the magnetic network5. In contrast to the dynamics, the differences in the appearance of emission lines and continua are minor, especially those from the layers of the low chromosphere6. It is therefore likely that the nonmagnetic and magnetic chromospheres are heated by the same mechanism. It is also noteworthy that the chromosphere produces emission everywhere and all the time4 and never the absorption spectrum predicted by the chromospheric bright point model of Carlsson and Stein7,8. The observations of chromospheric radiation thus suggest a separation of the description of chromospheric phenomena into (1) the dynamics of K2v bright points with 3-min oscillations of large amplitude at a few, select points in the internetwork medium, and the dynamics of network bright points with 7-min oscillations of large amplitude in magnetic flux tubes and (2) the general heating of the chromosphere. The paper is structured accordingly: Section 2 discusses chromospheric dynamics on the basis of the hydrodynamic and magnetohydrodynamic equations. Section 3 considers the general chromospheric heating requirements, while in 1. Introduction THE most prominent lines in the visible spectrum of the chromosphere are the H and K lines of Ca II. Their emission shows a separation of the atmosphere into nonmagnetic and magnetic regions, which are referred to as internetwork and magnetic network or, because of their connection to convection, as interior and boundary of supergranulation cells. This bifurcation of the medium is further emphasized by the dynamical behaviour, which shows oscillations with periods of 3 min in the internetwork chromosphere and 7 min in the network1 (Figure 1). Furthermore, the time variation of the intensity is different: The 3-min oscillations are marked by a sharply-peaked time variation of the intensity2 and a strong preference for the blue emission peaks, H2v and K2v, and the 7-min oscillations are accompanied by a broad time variation of the intensity and a much less pronounced † For correspondence. (e-mail: wolf@cfa.harvard.edu) 1496 Figure 1. Velocity power spectra from the Doppler motion of the H3 -absorption minimum of the H line, at disk center. Network and internetwork regions along the slit are averaged separately. From Lites et al.1 . CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Section 4 the generation of acoustic waves in the convection zone and the dissipation of this energy by weak acoustic shocks is outlined. Section 5 summarizes the paper. 2. Chromospheric dynamics Waves in the chromosphere are affected by the density stratification due to gravity, which makes waves dispersive, and introduces cutoffs that separate propagating from evanescent frequencies. For acoustic waves in the nonmagnetic medium, where gas pressure provides the restoring force in the wave equation, the cutoff frequency, νac , at the temperature minimum between photosphere and chromosphere is equal to 5 mHz (cutoff period Pac = 1/νac = 3 min). Waves with frequencies lower than νac are evanescent, i.e. their phase velocity is infinite and their group velocity is zero; thus they transport no energy (in the lowamplitude limit). For internal-gravity waves in the nonmagnetic medium, where gravity provides the restoring force in the wave equation, the cutoff frequency is the Brunt-Väisälä frequency, NBV. Waves with frequencies lower than NBV can propagate, but they are excluded from the purely vertical direction. In a neutral, monatomic gas, NBV and νac have practically the same value. Thus the two cutoff frequencies separate the wave spectrum into the regimes of low frequencies, where only internal-gravity waves can propagate, and high frequencies, where only acoustic waves can. In the magnetic chromosphere, two wave-modes play a role, namely, longitudinal flux tube waves, where gas pressure provides the main restoring force, and transverse flux tube waves, or kink waves, where the magnetic field does9. For the magneticfield strengths encountered in the quiet network, the cutoff frequencies are νλ = 5 mHz (Pλ = 3 min) for the longitudinal mode, and νκ = 2.5 mHz (Pκ = 7 min) for the kink mode. For both wave-modes, propagating waves have frequencies above the cutoff, and evanescent waves, below the cutoff. A dynamical model by Carlsson and Stein7 of the nonmagnetic chromosphere combined a sophisticated hydrodynamic code, incorporating NLTE radiative transfer of the Ca II ion, with empirical driving, which was taken from the Doppler velocity measured in a photospheric Fe-I line in an hour-long observing run1. The empirical approach sacrificed the search for the underlying cause of the oscillations, but, in return, allowed firm conclusions about the nature of the K2v phenomenon. Since the simulation7 reproduced to high fidelity the intricate intensity and velocity variations in the core of the H line for two out of the four bright points from the same observing run, it is clear that the waves powering K2v-bright points are propagating acoustic waves. In the network bright points on the supergranulation cell boundary, the periods of oscillation, typically near 7 min, are longward of the acoustic cutoff period and of the Brunt-Väisälä period, where acoustic waves are evanescent but internal-gravity waves propagate. Some observers have therefore considered the waves in the magnetic network to be internal-gravity waves10,11. A theoretical underpinning of this CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 hypothesis by Lou12 found the observed periods to be possible as resonances of magneto-gravity waves in a chromospheric cavity. But a heuristic picture of network bright points in terms of internal-gravity waves by Deubner and Fleck13 was shown by Kalkofen14 to fail in its intended purpose, namely, to dissipate the wave energy in traveling downward in the narrowing flux-tube funnels. An explanation of network bright points not based on oscillations in a cavity was proposed by Kalkofen14 who noted the coincidence of the observed oscillation period with the cutoff period of kink waves for magnetic field strengths found in the magnetic network. Numerical modeling by Choudhuri et al.15 and the analytic solution of the MHD equations for impulsive excitation16 showed that oscillations at the cutoff could indeed be produced. In order to dissipate, however, the transverse waves have to transfer energy to the longitudinal mode, which forms shocks17. Wave excitation in a magnetic flux tube produces both transverse and longitudinal waves. But the power spectrum of waves in the network shows virtually no power at 5 mHz, the cutoff frequency of the longitudinal mode. The scenario for network-bright points therefore requires the wave excitation to produce mainly transverse waves. This is indeed the case, as was shown with numerical solutions18 and with analytic methods19. A consequence of the almost exclusive excitation of transverse waves is that a significant flux in longitudinal waves appears only in the chromosphere, where the nonlinearity of the waves facilitates the transfer of energy between modes. This transfer is consistent with the low value of the observed velocity coherence between the base and the middle of the network chromosphere1 (Figure 7), which indicates that the dissipating longitudinal waves do not arrive from below but are formed in the chromosphere. For the analytic modeling of chromospheric oscillations we consider the hydrodynamic equations. From a small-amplitude expansion of the equations for a onedimensional, isothermal atmosphere we obtain the Klein–Gordon equation20: ∂2 ∂z 2 u− ∂2 ∂t 2 u − u = 0, (1) which is written here for the ‘reduced’ velocity u and in terms of the dimensionless depth and time variables z and t. For acoustic waves, the ‘physical’ velocity v is obtained as v (ζ, τ) = µ(ζ, τ)eζ /2H , (2) in terms of the physical depth and time variables ζ and τ. The derivation of the wave equation (1) for acoustic waves assumes that the atmosphere is one-dimensional, isothermal and stratified in plane-parallel layers with a constant scale height H. In the vertical direction, only acoustic waves are allowed (i.e. internal-gravity waves are excluded from the vertical direction). The velocity v grows exponentially with height ζ with a scale length of twice the density-scale height. 1497 SPECIAL SECTION: SOLAR PHYSICS Magnetic waves propagating in the expanding geometry of a thin magnetic flux tube, in pressure equilibrium with the surrounding medium, satisfy the same wave equation20, but with different definitions of the dimensionless variables. An additional requirement (which is satisfied here) is a constant ratio β(= 8πp/B2) of gas to magnetic pressure inside the tube. The physical velocity v is now obtained as v (ζ, τ) = u(ζ, τ)eζ /4H, (3) showing slower exponential growth for tube modes, with a scale length of 4H because of exponential spreading of the tube cross section. The solution of the wave equation for a velocity impulse at z = 0 and t = 0, in an infinite medium is given by Lamb21, u ( z, t ) = 12 δ (t − | z |) − 2 2 t J1 ( t − z ) 2 t 2 − z2 H ( t− | z |), (4) where δ is the Dirac δ-function; H is the Heaviside function and J1 is a Bessel function. The argument of both the δ-function and the Heaviside function expresses the propagation of the head of the wave (|z| = t), at the sound speed for acoustic waves and at the tube speeds for the flux-tube waves. The tube speeds in dimensioned units are given by cλ = cs [2/(2 + γβ)]1/2 for the longitudinal mode, and cκ = cs [2/(γ(1 + 2β))]1/2 for the transverse mode, where cs is the sound speed and γ (= 5/3) is the ratio of specific heats. Behind the head of the wave, the atmosphere oscillates initially with frequencies in a broad spectrum, but this narrows with time until it results in an oscillation at the cutoff 22, as shown by the asymptotic solution of the Klein– Gordon equation: u ( z, t) ~ cos (t ) t , t >> z ≥ 0 , (5) in which the (reduced) velocity becomes independent of height z. The solution implies that gas elements at all heights reach maximal amplitude simultaneously. At the cutoff22 frequency, the phase velocity, which is given by ω , v group = 1/ vφ , ω2 − 1 is infinite and the group velocity is zero. In dimensioned units, the cutoff periods are given by vφ = (6) Pac = 4πH/c s , Pλ = Pac ( 60 + 50 β) /(63 + 48 β ) , (7) Pκ = Pac 2γ (1 + 2 β) , for acoustic waves and for longitudinal and transverse flux-tube waves, respectively22. It is interesting to note that the highest phase velocities are 1498 reached near the origin of the wave, and that they increase with the order of the maximum behind the head of the wave16. For comparison with observations it needs to be remembered that these analytic results are for an isothermal atmosphere initially at rest; even the observed values of the photospheric velocity allowed to reproduce the H-line observations only when the waves were launched into a disturbed atmosphere7. The asymptotic solution of the Klein–Gordon equation shows that an impulse can excite oscillations at the cutoff of the respective mode. Any sudden change in the forcing function has the same consequence. An example of the former is found in the seismic events following the collapse of an intergranular lane23, an example of the latter in stochastic excitation24. Thus there are two possible models for the generation of K2vbright points: (1) Individual events at a few, discrete points in the internetwork medium. They might account for the strong bright points observed by Liu (1974)25 or by Brandt et al. (1992)26 with a filling factor of about 1%, or modeled7. (2) Ubiquitous generation of oscillations from the turbulence in the convection zone27, leading to bright points that are visible anywhere in the K line2, or in UV lines4 where they are observed in half the positions along the slit of the SUMER instrument. For network bright points Muller et al.26 have observed the interaction of fast granules with magnetic-flux tubes. This process was modeled by Choudhuri et al.15, as footpoint motion of a flux tube, resulting in a kink wave in the tube. While the various processes make plausible the excitation of oscillations at the cutoff of a mode, observational confirmation of a link between a photospheric event and, after a delay accounting for the wave travel time, of a chromospheric brightening is still lacking. An interesting puzzle is posed by the dynamical modeling of H2v-bright points7. On one hand, the simulation reproduced to high accuracy the complex variation of the intensity and shape of the H line, including the proper time delay between the motion of the photospheric Fe-I line and that of the resulting H line, and on the other hand, it predicted an absorption spectrum from any location in the internetwork chromosphere and most of the time, whereas the observations4 with SUMER show only an emission spectrum. The high spatial and temporal resolution of the space observations exclude the possibility of spatial and temporal averaging to hide a chromospheric absorption spectrum behind the observed emission. Thus the chromosphere must have a permanent temperature rise in the outward direction, as shown by empirical models. The flaw in the model7 was traced to the input velocity spectrum by Kalkofen et al.28 who argue that the dynamical model uses only about 1% of the acoustic energy flux available. The hidden flux has frequencies in excess of 10 mHz. The flux critical for the bright point phenomenon is emitted between 5 mHz and 10 mHz. That flux is found in the observed velocity spectrum and accounts for the success of the dynamical simulation. In the following section we address the generation of the flux that is missing from the simulation7 and treat its dissipation by means of the theory of weak shock waves. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Chromospheric energy losses and the heating requirements The empirical chromosphere model of Vernazza et al.6 has been constructed by fitting the observed ultraviolet C I, Si I and H I continua with the simulated emission from a hydrostatic equilibrium model with an arbitrary temperature distribution. This temperature distribution was subsequently modified until an optimal fit between the observed and simulated emission was obtained. For the average sun model C, these authors subsequently computed the chromospheric energy-loss rate in H, H–, Ca II and Mg II due to lines and continua. They found a total radiative energy loss of FR = 4.6 × 106 erg cm–2 s–1. Anderson and Athay29 later improved on this determination by including the abundant line emission from Fe II and found a total chromospheric radiative energy loss of FR = 1.4 × 107 erg cm–2 s–1. In addition, they found (see Figure 2) that the cooling rate ΦR in most of the chromosphere is proportional to the mass density ρ0, which leads to a characteristic height dependence of the chromospheric emission. The question is how this persistent energy loss is balanced and how the continuous supply of energy is provided. The time scale in which an excess temperature would cool down to the boundary temperature if the mechanical heating were suddenly disrupted is given by the radiative relaxation time for which one has t Rad = ∆E ρcv ∆T ρcv = = ≈ 1 .1 × 10 3 s. Φ R 16κσT 3 ∆T 16 κσT 3 (8) Here we used from the VAL81 model that at z = 1280 km, T = 6200 K, ρ = 9.8 dyn/cm2, κ /ρ = 4.1 × 10–4 cm2/g, cv = 9.6 × 10–12 g/cm3, σ = 5.6 × 10–5 erg/cm2 s K4. It is seen that in timescales of a fraction of an hour, the chromosphere would cool down to the boundary temperature if mechanical heating would suddenly stop. In a purely hydrodynamic environment there are few possibilities to persistently heat a medium: acoustic waves or pulsational waves. For a review on heating mechanisms see Narain and Ulmschneider30. Acoustic waves generated in the convection zone are the most likely possibility. To see what acoustic wave heating might do, consider a typical acoustic disturbance in the solar chromosphere. Assume a characteristic perturbation of size L = 200 km, temperature ∆T = 1000 K and velocity ∆v = 3 km/s. Using appropriate values for the thermal conductivity κth = 105 erg/cm s K and viscosity ηvis = 5 × 10–4 dyn s/cm2 we find for the thermal conductive and viscous heating rates ΦC = d dT κ th ∆T κ th ≈ ≈ 3 × 10 − 7 dz dz L2 2 ΦV erg 3 , cm s dv n ∆v 2 = n vis ≈ vis 2 ≈ 1 × 10 − 7 L dz erg 3 . cm s CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 These heating are inadequate by a large margin to bal-ance the typical empirical chromospheric cooling rate of 10–1 erg/cm3 s found in the model6. Only when the length scale L is decreased by several orders of magnitude can the heating rates be raised to acceptable levels. For acoustic waves, this is accomplished by shock formation. This shows that in pure hydrodynamic situations only shock heating has sufficient power to balance the observed cooling rates. The same very likely is the case for the heating of chromospheric magnetic flux tubes, where the shock heating is by longitudinal tube waves. Note as discussed in the review30, this is different in the transition layer where in addition other types of heating, like microflare heating become important. 4. Weak shock heating and the generation of acoustic waves Small-amplitude acoustic shock waves behave essentially like acoustic waves, except that they dissipate at the shock front (see below). In particular, the amplitude relations are identical. Consider linear small-amplitude sawtooth waves with pressure and velocity variations p = p0 + pm – 2p mt/P, v = v m – 2v mt/P, where P is the wave period, t the time and subscript m indicates maximum amplitudes, the wave energy flux (erg cm–2 s–1) is given by: FM = 1 P 1 1 ∫0 ( p − p0 ) v dt = 3 pmv m ≈ 12 γ p0 cSη , P 2 (11) where p 0 is the unperturbed pressure, cS the sound speed, γ the ratio of specific heats and where for weak shocks one has for the total pressure, velocity, temperature and density jumps (9) lim (10) Figure 2. Comparison of the theoretical limiting acoustic flux FM (erg cm–2 s–1 ) with the solar chromospheric radiative loss flux Fr determined empirically by Anderson and Athay29 . ΦR (erg cm–3 s–1 ) is the empirical net radiative cooling rate. PA/5 and PA/10 label different assumptions as to the acoustic frequency spectrum. Star symbols show empirically determined acoustic fluxes by Deubner39 . 1499 SPECIAL SECTION: SOLAR PHYSICS 2p m ≈ γp 0η, 2v m ≈ cSη, 2Tm ≈ (γ – 1)T0η, 2ρm ≈ ρ0η. Here the shock strength is defined as η = (ρ2 – ρ1)/ρ1, where ρ1, ρ2 are the densities in front and behind the shock31. By expanding the entropy jump ∆S per unit mass at the shock front, the shock dissipation rate (erg cm–3 s–1) of the wave can be written ΦMM = ≈ ρT ∆S P = p ρ −γ ln 2 2 γ (γ − 1) P p 1 ρ1 ρ0 c S2 1 γ (γ + 1) p 0 η3 . 12 P (12) The approximate equality is only valid for weak shocks where the entropy jump is small. Let us assume a gravitational atmosphere and in analogy to ray optics that the quantity is conserved. FM cS2eq. (11) with respect to height z and using (12) Differentiating gives an equation for the shock strength 2 dη η γ g 3 dcS (γ + 1)η = − − , dz 2 cS2 cS P 2cS2 dz term of eq. (13). Limiting strength is reached when η becomes constant and the flux proportional to the gas pressure p 0. Figure 2 shows limiting-strength acoustic-heating fluxes lim for wave F periods P = PA/5 and P = PA/10, where PA = 4πcS/(γg) is M the acoustic cut-off period. It is seen that the fluxes log rise linearly with the logarithm of the FMlimmass column density m, since m = p 0/g and from eq. (15) F lim ~ p 0. The figure therefore shows that acoustic shock waves are able to explain the observed height dependence of the chromospheric emission. As will be shown below, acoustic wave generation calculations for the solar convection zone by Musielak et al.32 also provide the correct wave period for Figure 2. As shown below one finds that P = PA/5 is roughly the period of the maximum of the acoustic-wave spectrum generated in the convection zone. Using this value one has from the above equations ηlim = 1 4π = 0 .94 , 5 γ +1 (13) FMlim = where g is the gravitational acceleration. The refractive term dcisS2 /small dz in the chromosphere and can be involving neglected. For an isothermal, nonionizing, gravitational atmosphere eq. (13) can be solved for various initial shock strength’s η0 and shows that irrespective of the η0 the shocks eventually reach a limiting shock strength ηlim given by γ gP ηlim = . (14) (γ + 1)cS 4π 2 γ c S p 0 ≈ 0.123 cS p 0 75 (γ + 1) 2 1 γ3 g2 P 2 p0 . 12 (γ + 1) 2 cS and a heating rate per gram 1 dFM ρ dz = dFM lim dm = ΦM ρ = γ 4π 2 cS g 75 (γ + 1) 2 (18) 9 erg gg–1-1 ss–1-1. . ≈≈22.4 .4 ⋅×10109 erg (15) Eq. (13) has the property that for small shock strengths one initially has an exponential growth due to flux conservation which results from the first term on the RHS of the equation. This growth of the sawtooth wave is similar to that for acoustic waves in a gravitational atmos2 2 phere assuming flux conservation, ρ0v 2~ ρ 0 cS η = const. The increase in shock strength is eventually balanced by the increasing shock dissipation described by the last 1500 (17) ≈ 10 5 p 0 ≈ 32..74m ×10 , 9 m, From eq. (11) one obtains a limiting wave energy flux FMlim = (16) The computation of the generation of acoustic energy in stellar convection zones dates to the theory of quadrupole sound generation from turbulence developed by Lighthill33,34 and Proudman35. Considering a Kolmogorov turbulence spectrum, Lighthill has derived an expression FM = ∫ 38 ρ0 u 8 cS5 H dz, (19) CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Figure 3. Acoustic wave spectra for different values of the mixing-length parameter α (left panel), different contributions to the acoustic spectrum (right panel), computed with the Lighthill–Stein theory, after Musielak et al.32 . called Lighthill formula, where z is height, H is the scale height, ρ0 the density, cS the sound speed and u the turbulent velocity. This famous u 8-formula was found to be in excellent agreement with measurements in terrestrial applications. Lighthill’s theory was further extended by Stein36,37 to allow for the computation of the acoustic frequency spectrum and was recently revisited32. To compute the acoustic flux using the Lighthill–Stein theory, one must first calculate a convection zone model where one uses the mixinglength theory, which depends on a free parameter, the mixinglength parameter α. Musielak et al.32 on the basis of a description of the turbulence with an extended Kolmogorov spectrum and a modified Gaussian frequency factor, found total acoustic fluxes FM = 1.3 × 107 erg cm–2 s–1 for α= 1.0 and 1.7 × 108 erg cm–2 s–1 for α= 2.0. Figure 3 shows the acoustic spectra obtained by these authors and demonstrates that quadrupole generation (also used in the Lighthill formula) is the most important contribution to the acoustic-wave flux. Since the frequencies in Figure 3 are circular frequencies, the spectra have a maximum at periods P = 79, 58 and 41 s for α= 1.0, 1.5, 2.0, respectively. Recent numerical convection zone calculations show that α, the ratio of the mixinglength L to the scale height H, typically varies only in a narrow range of α≈ 2.0–2.16 (ref. 38). Taking a temperature T = Teff = 5770 K, the acoustic cut-off period is found to be 216 s such that P = PA/5 = 43 s is a good estimate for the maximum of the generated solar acoustic wave spectrum. Note that Figure 2 shows also empirical acoustic fluxes by Deubner39 which are in rough agreement with the empirical losses. 5. Conclusions Waves are responsible for the characteristic features of the chromosphere: the permanent outward temperature rise, which is due to acoustic waves that dissipate in shocks, the 3-min oscillations of the nonmagnetic chromosphere, whose period matches the cutoff period of acoustic waves, and the 7-min oscillations, whose period matches the cutoff period of transverse CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 flux tube waves for a pressure ratio of β ≈ 0.25. The empirical radiative emission rate of the chromosphere can be understood on the basis of the dissipation rate of weak-shock theory, and the appearance of the cutoff periods in the chromospheric oscillations can be understood as due to impulsive or stochastic wave excitation, which causes the atmosphere behind the head of a wave to oscillate at the cutoff period. 1. Lites, B. W., Rutten, R. J. and Kalkofen, W., Astrophys. J., 1993, 414, 345. 2. von Uexküll, M. and Kneer, F., Astron. Astrophys., 1995, 294, 252. 3. Grossmann-Doerth, U., Kneer, F. and von Uexküll, M., Solar Phys., 1974, 37, 85. 4. Carlsson, M., Judge, P. G. and Wilhelm, K., Astrophys. J., 1997, 486, L63. 5. Curdt, W. and Heinzel, P., Astrophys. J., 1998, 503, L95. 6. Vernazza, J. E., Avrett, E. H. and Loeser, R., Astrophys. J. (Supplement), 1981, 45, 635. 7. Carlsson, M. and Stein, R. F., in Chromospheric Dynamics, Proceedings of the Mini-Workshop, Institute for Theoretical Astrophysics, Oslo, 1994. 8. Carlsson, M. and Stein, R. F., in IAU Symposium 185 (eds Deubner, F.-L., Christensen-Dalsgaard, J. and Kurtz, D.), Kluwer, Dordrecht, 1998, 435. 9. Spruit, H. C., in The Sun as a Star (ed. Jordan, S.), NASA Monograph Series on Nonthermal Phenomena in Stellar Atmospheres, 1981, p. 385. 10. Damé, L., Thèse, Université de Paris, VII, 1983. 11. Damé, L., in Small-Scale Dynamical Processes in Quiet Stellar Chromospheres (ed. Keil S. L.), NSO/SPO, Sunspot, NM, 1984, p. 54. 12. Lou, Y.-Q., MNRAS, 1995, 276, 769. 13. Deubner, F.-L. and Fleck, B., Astron. Astrophys., 1990, 228, 506. 14. Kalkofen, W., Astrophys. J., 1997, 486, L145. 15. Choudhuri, A. R., Auffret, H. and Priest, E. R., Solar Phys., 1993, 143, 49. 16. Kalkofen, W., Rossi, P., Bodo, G. and Massaglia, S., Astron. Astrophys., 1994, 284, 976. 17. Zhugzhda, Y. D., Bromm, V. and Ulmschneider, P., Astron. Astrophys., 1995, 300, 302. 18. Ulmschneider, P., Zähringer, K. and Musielak, Z. E., Astron. Astrophys., 1991, 241, 625. 1501 SPECIAL SECTION: SOLAR PHYSICS 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. Hasan, S. S. and Kalkofen, W., Astrophys. J., 1999, (in press). Rae, I. C. and Roberts, B., Astrophys. J., 1982, 256, 761. Lamb, H., Proc. Lond. Math. Soc., 1909, 7, 122. Spruit, H. C. and Roberts, B., Nature, 1983, 304, 401. Goode, P. R., Strous, L. H., Rimmele, T. R. and Stebbins, R. T., Astrophys. J., 1998, 495, L27. Sutmann, G. and Ulmschneider, P. Astron. Astrophys., 1995, 294, 232. Liu, S.-Y., Astrophys. J., 1974, 189, 359. Muller, R. and Roudier, Th., Vigneau, J. and Auffret, H., Astron. Astrophys., 1994, 283, 232. Theurer, J., Ulmschneider, P. and Kalkofen, W. Astron. Astrophys., 1997, 324, 717. Kalkofen, W., Ulmschneider, P. and Avrett, E. H., Astrophys. J. (Lett.), 1999, 521, (in press). Anderson and Athay, Astrophys. J., 1989, 346, 1010. Narain, U. and Ulmschneider, P., Space Sci. Rev., 1996, 75, 453. Ulmschneider, P., Solar Phys., 1970, 12, 403. 1502 32. Musielak, Z. E., Rosner, R., Stein, R. F. and Ulmschneider, P., Astrophys. J., 1994, 423, 474. 33. Lighthill, M. J., Proc. R. Soc. London, 1952, A211, 564. 34. Lighthill, M. J., Proc. R. Soc. London, 1954, A222, 1. 35. Proudmann, I., Proc. R. Soc. London, 1952, A214, 119. 36. Stein, R. F., Solar Phys., 1967, 2, 385. 37. Stein, R. F., Astrophys. J., 1968, 154, 297. 38. Trampedach, R., Christensen-Dalsgaard, J., Nordlund, A., Stein, R. F., in Solar Convection and Oscillations and their Relationship (eds Pijpers, F. P., Christensen-Dalsgaard, J., Rosenthal, C. S.), 1997, p. 73. 39. Deubner, F.-L., in Pulsation and Mass Loss in Stars (eds Stalio, R., Willson, L. A.), Kluwer, 1988, p. 163. ACKNOWLEDGEMENTS. W.K. thanks his colleagues at the Institut für Theoretische Astrophysik for their hospitality and the University of Heidelberg for a Mercator guest professorship funded by the DFG. Partial support by NASA is acknowledged. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Solar activity: An overview v Zdenek Š vestka CASS UCSD, La Jolla, California, CA 92093-0424, USA and SRON Laboratory for Space Research, Sorbonnelaan 2, 3584 CA, Utrecht, The Netherlands This review article describes briefly the main characteristics of the active Sun: the different active phenomena, the 11-year cycle of their appearance, and their influence on the environment of our Earth. 1. Solar cycles Although the Sun illuminating our Earth looks like a steadily shining celestial body, its surface is actually the seat of continuous changes and powerful activity. As the Solar and Heliospheric Observatory (SOHO) spacecraft recently revealed, in ultraviolet lines the solar surface looks like ‘boiling’ all the time and everywhere one can see variations in brightness, plasma flows, and small ejections of gas, indicating permanent changes of the structures in the solar atmosphere. But this is not what we call solar activity – all these changes are still considered to occur on the ‘quiet Sun’. The real processes, called solar activity, which have their impacts also on the Earth environment, appear in limited parts of the solar atmosphere, and their occurrence varies quasi-periodically with time, creating 11-year cycles of solar activity. Each new solar cycle is born close to the solar poles and its activity then slowly propagates to lower heliographic latitudes. The real length of one cycle is actually about 22 years, but only the second half of it begins to produce clearly visible active processes on the Sun. When the Sun is viewed in white light, one observes the lowest level of the solar atmosphere, which is called the photosphere. In the photosphere, solar activity manifests itself as sunspots or groups of sunspots (Figures 1 and 2 a). Therefore, solar cycles were long characterized (and still are) by the so-called relative sunspot numbers R: A daily R is the sum of the number of sunspot groups on the Sun plus the number of individual sunspots in all of them. A monthly or yearly R is the average daily R during a month or a year. Individual solar cycles differ in their lengths and heights – lengths varying between 9 and 16 years, and yearly maximum R varying between 46 and 190 have been observed between the half of the 18th century and present days. Rare sunspot groups begin to appear first at high latitudes (between 40 and 50 heliographic degrees) and as the frequency of their occurrence increases, their positions e-mail: z. svestka@sron.nl CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 move progressively closer to the equator. A few years after the onset of an active cycle, when the spots appear mostly at latitudes below 25°, the cycle reaches its maximum. Thereafter it slowly declines, with the spot occurrence approaching the equator, and eventually reaches a minimum when sunspots appearance becomes very rare and for many days the Sun is without any sunspots. However, usually before the last sunspots of the old cycle disappear, new ones begin to appear at high latitudes. 2. Active regions This solar activity is due to magnetic field which exists in the Sun, generated by electric currents. Sunspots become visible in the photosphere when ropes of magnetic flux emerge on the solar surface. The magnetic flux in the central dark umbra of a spot is usually between 1500 and 3000 gauss and, as a consequence of this strong field, temperature in a spot is much lower (4000 K or less) than in the surrounding photosphere. This causes the dark appearance of sunspots in which the umbra is surrounded Figure 1. A sunspot group in the photosphere (above) and the photospheric magnetic field (below). Black and white denote the opposite magnetic polarities. 1503 SPECIAL SECTION: SOLAR PHYSICS by a less dark penumbra (Figure 1). However, as Figure 2 b shows, the situation looks quite differently if we observe sunspot groups in the higher layer of the solar atmosphere, the chromosphere. Since about 1930, this layer could be observed in a spectrohelioscope and through filters, which make it possible to observe the Sun in a narrow monochromatic band, usually centered on the hydrogen-Balmer-Hα line. In this line, sunspots are surrounded by bright plages, and we call these groups of plages and spots the active regions on the Sun. And since 1973, when Skylab orbited the Earth and imaged the Sun in X-rays, we can also see the highest layer of the solar atmosphere, the solar corona in which these active regions appear extensive and bright (Figure 2 c). Active regions appear where magnetic flux emerges from subphotospheric layers to the solar chromosphere and corona (see the magnetic map in Figure 1). There are many flux emergences on the solar surface, all the time creating ephemeral active regions which in X-rays are seen as bright points on the Sun (examples – bright dots –can be seen in Figure 2 c). However, only few of them grow further, with newly emerging flux continuously added to them and eventually developing into much larger active regions (four of them are seen in Figures 2 a, b and c and many more on the full-disk picture of the Sun in the Hαline in Figure 3). Most active regions are bipolar, with two main spots and surrounding plages of opposite magnetic polarities. The two polarities are connected by chromospheric and coronal loops, particularly well seen in X-rays and ultraviolet lines (Figure 4), and are separated by a neutral line below the tops of these loops where the longitudinal magnetic field is zero. (Compare the schematic drawings in Figures 9 a and b.) Solar atmospheric gas slowly accumulates along these lines and cools. In the Hαline one can see them along these neutral lines dark filaments (many are seen in Figure 3) which on the solar limb look like bright prominences (Figure 5). Dark filaments often survive even after the active region decays, and can form outside of active regions as well (like that one to the southwest of the Sun’s center in Figure 3). In some cases, irregular patches of magnetic flux emerge in an active region and cause irregularities in its magnetic structure. Then several different neutral lines exist in the region and more dark filaments can form there. While all active regions are seats of various kinds of active processes, the most powerful events of solar activity occur in these magnetically complex active regions, where both magnetic polarities are mixed. Most active is the so-called δ configuration, when umbrae of opposite polarities are embedded in one common penumbra. 3. Complexes of activity and interconnecting loops Figure 2. Images of four active regions in (a) white light, (b) Hα line, and (c) X-rays. 1504 Solar activity seems to prefer, sometimes for a period of many months, selected active longitudes where active regions form more frequently than elsewhere on the Sun, creating there so-called complexes of activity (Figure 6). This seems to be due to some irregularities in distribution of the subphotospheric magnetic flux that emerges through the photosphere. Once such an irregularity is formed, it takes a long time before effects of the solar rotation, which slightly varies with the latitude, remove it. Many active regions in such a complex of activity are CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS connected by coronal loops that very often extend across the equator. We call them interconnecting loops. They can be best observed in soft X-rays, as was first done on Skylab in 1973 (an example is shown in Figure 2 c) and presently by Yohkoh. Due to magnetic field variations in the interconnected active regions, these connections vary all the time in their shape and brightness. In the new solar cycle, when active regions begin to emerge at high latitudes, these interconnections have been found very long, up to 60 heliographic degrees (700,000 km, i.e. 55 diameters of the Earth, Figure 7). It is quite impossible that such long connections could possibly exist below the photosphere and emerge through it into the corona. The most plausible explanation is that they are formed by magnetic field-line reconnection in the solar corona. (Figure 8 b) which have some features common with surges, but apparently are not exactly the same phenomena. Most probably, as the new satellite TRACE recently revealed, both have a similar cause, but the hot jets and cold surges move along different trajectories through the corona. In the past few years, the sophisticated spacecraft SOHO and TRACE could detect some jets even in the quiet Sun regions, reflecting the complexity of magnetic field structure even outside the active regions. A much more powerful phenomenon seen in Hαimages is a spray (Figure 8 c), in which large amounts of active region plasma are abruptly ejected into the corona, and often escape into interplanetary space, possibly being one of the sources of coronal mass ejections (cf., Section 6). 4. Surges, jets, and sprays In most active regions, variations in brightness occur all the time reflecting either new emergence of magnetic flux or oldflux decay or interactions between individual active region loops. Small short-lived brightenings in active regions have been called Ellerman bombs or moustages. In the magnetically complex regions also ejection of material occurs, mostly rooted inside tiny patches of magnetic Figure 4. Image of the active Sun in soft X-rays (made on Yohkoh). Note the sets of bright loops that cross the neutral lines in active regions. polarities embedded inside, or penetrating into, a region of opposite polarity which the moustages prefer. In the Hα line the most common ejection is a bright or dark surge (Figure 8 a) in which solar material first moves upward into the corona along magnetic field lines and subsequently falls back to the chromosphere. In X-rays, the Japanese satellite Yohkoh observed many bright jets CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 Figure 6. High-resolution photograph of a complex of activity near the western solar limb. Some sunspots and many dark filaments can be seen and new magnetic flux is emerging in the middle Figure 5. and Prominence observed the See limbalso in the (Big foreground at the limb (upper on right). the Hα fine line structure Bear It would be seen filament in of theObservatory surroundingphotograph). ‘quiet’ chromosphere. Northas isa dark to the left and projection the(Hα solar disk. west to theon top. line, Ottawa River Solar Observatory.) 1505 SPECIAL SECTION: SOLAR PHYSICS 5. Solar flares However, the most powerful brightening in an active region is a solar flare. In the optical range almost all flares can be seen only in monochromatic light (most observations are made in the Hα line), but the most powerful ones also emit in the white light and the first flare ever detected was discovered by Carrington on 1 September 1859 when he observed a large sunspot group looking at the photosphere. After Hale’s invention of the spectrohelioscope, which made it possible to observe continuously the whole solar surface in the Hα line, from early thirties flares have been observed regularly at many solar observatories throughout the world and listed in monthly reports. As the resolving power of solar instruments was improving, smaller and smaller flares and flare-like phenomena could be detected in active regions on the Sun. Thus first the category of subflares has been added to the original flares, and later the categories of microflares and still smaller nanoflares. Obviously, flare-like processes can be detected in an active region on all scales and an idea – originally due to Parker – is that nanoflares occur everywhere on the Sun and are the actual source of coronal heating. Generally, flares are of two different kinds: Confined flares, where preexisting loops in an active region suddenly brighten and thereafter slowly decay; and eruptive flares, where the whole configuration of loops crossing a neutral line in an active region is disrupted and must be newly rebuilt. Most flares, and essentially all small flares are confined flares, and quite often they originate through an interaction of two active region loops which magnetically reconnect. But the most powerful and energetic phenomena are the eruptive flares which are also one of the sources of a b c d Figure 9. The interpretation of eruptive flares. a and b, Two different views of the preflare situation when a dark filament (prominence) extends along a neutral line and is embedded in a system of loops forming a coronal helmet structure; c, Opening of magnetic field lines; d, Subsequent closing of field lines, creating the flare loops. 1506 coronal mass ejections (cf. section 6). Figure 9 shows schematically the development of an eruptive flare, following the original suggestion of Kopp and Pneuman in 1976, later improved and further developed by many other authors. The originally closed magnetic field Figure 7. The longest transequatorial interconnecting loop of the new solar cycle connecting active regions over the distance of 700,000 km. (Yohkoh soft X-ray image of the Sun.) Figure 8. a. Bright surge in the Hα line, observed at Big Bear Solar Observatory; b. X-ray jet observed by Yohkoh; and c. a spray, observed at Wroclaw Observatory. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS in an active region, in which a filament (prominence) is embedded, suddenly opens. Reasons for it can be a newly emerging magnetic flux, a confined flare nearby, a wave disturbance coming along the solar surface from another source of activity, or some internal instability (e.g. a shear of field lines which exceeds a certain critical limit). As field lines open (Figure 9 c) plasma begins to flow from the dense chromosphere upward to the corona, so that gas pressure decreases and magnetic pressure begins to prevail. That leads to sequential reconnections of the open field lines, which begin to create new loops in the active region (Figure 9 d). The reconnection process produces intense heating at the top of each new loop which is conducted downward to the chromosphere, and it also accelerates particles which flow along the loop to its footpoints. Thus the gas at the chromospheric footpoint is strongly heated and evaporates into the newly formed loop, making it visible in X-rays and high-temperature lines as a flare loop. The loop then cools, but in between other loops are formed above it through reconnection of other field lines and the whole process is repeated in each of them. Thus the loop system gradually grows. After some time, the lowest loops cool to about 10000 K and begin to be visible in the Hα line. This takes some time, so that earlier, when no X-ray observations of eruptive flares were available, these structures were called post-flare loops. While in compact flares energy is suddenly released and thereafter the flare structure cools and decays, in eruptive flares energy is released during each reconnection and thus this process of energy release, though decreasing in efficiency, can continue for many hours. Because each reconnection produces new X-ray flux, one can see Figure 10. Composed image of an eruptive flare near the solar limb. A quiescent prominence, like that in Figure 5, can be seen in the background. (Images made and composed at Wroclaw Observatory.) CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 enhanced emission in X-rays during the whole time of the repeated reconnections. Therefore, based on these X-ray records, eruptive flares are also often called long-duration events. When looking at the chromospheric image of an eruptive flare in the Hαline, one observes at the footpoints of newly formed coronal loops the heated chromosphere, in the form of two bright ribbons, which slowly separate as the flare loop system grows. Therefore, these phenomena were earlier called (and often still are) two-ribbon flares. The two ribbons are connected by bright or dark loops. Figure 3 shows an example of such a flare (to the southwest of the center) and the bright patches between the bright ribbons are tops of the loops that connect them. And a more detailed photograph of an eruptive flare, composed from three different images, is shown in Figure 10: we see there the sunspot group in the photosphere, the two bright ribbons in the chromosphere, and the ‘post’-flare loops seen in the Hα line above the limb. Both in the Hα line and in X-rays one can see how in all eruptive flares the loop system, while decaying, slowly grows, sometimes for many hours. 6. Coronal mass ejections The opening of magnetic field lines which initiates an eruptive flare is connected with ejection of material. This was first recognized some 30 years ago on metric radio waves, when strong outbursts of radio emission were seen high in the corona, some continuously moving upward. But only spacecraft observations (first by Skylab in 1973) showed plasma ejections from the Sun, now called coronal mass ejections (CMEs), which moved with high speeds into interplanetary space. (See examples in Figure 11.) Eruptive flares are one of the sources of these CMEs which, as we know now, play the most important role in solar-terrestrial relations. But eruptive flares are not the only source of the CMEs. Everywhere on the Sun magnetic field lines, closed across a neutral line, can be disrupted, open, and eject solar plasma into space. Eruptive flares are only one special – and apparently the most energetic – case of the field-line opening when strong magnetic field inside an active region is involved in the process. Because in many cases the neutral line inside an active region is marked in Hα line images by a dark filament, the opening of magnetic field is often first made apparent by its activation. The filament structure begins to change, parts of the filament slowly rise into the corona, the speed of the rise accelerates, and finally the whole filament erupts. Only later on, when the open field lines begin to reconnect, bright flare loops begin to appear, but at that time the ejected material (often with the filament embedded in it) already propagates with a speed of a few hundred km/s high in the corona into interplanetary space. 1507 SPECIAL SECTION: SOLAR PHYSICS We can often see a very similar process far from any active region. A quiescent filament (several are seen in Figure 3) becomes activated and eventually erupts. Only, on the quiet Sun, we do not see in Hα any eruptive flare, because magnetic field there is not strong enough to produce all the flare effects which we see in active regions. These activations of quiescent filaments have been known for some 50 years and called disparition brusques, but only in the seventies space observations in X-rays revealed that these disruptions can have equally important effects in interplanetary space and at the Earth as major flares in active regions. However, field lines crossing neutral lines can open also at places where no dark filament exists, so that no eruption at all is visible in the Hα line. In X-rays or in spectral lines corresponding to high temperatures, all these field openings have some detectable effect, but such observations are not carried out so often as in the Hα line, so that these responses can easily be missed. This fact is one of the reasons why for a long time the real sources of many CMEs remained unknown. Another reason probably is that not all CMEs originate in this way. Some may be connected with ejections of material along magnetic field lines, for example in sprays (cf., Section 4). intense events deep into interplanetary space, up to the Earth distance. And particles trapped in magnetic clouds associated with CMEs produce powerful Type IV bursts, partly stationary and partly moving upward. Accelerated electrons also give rise to hard X-rays in solar active regions by bremsstrahlung, and particles accelerated to high energies produce in some flares γ-rays which are partly continuum, originating through bremsstrahlung of relativistic electrons, and partly lines excited through electron–positron annihilation, neutron capture by protons and helium nuclei in the photosphere, or by transitions in excited nuclei of heavier atoms. Particles propagating through interplanetary space produce disturbances in the Earth environment, about which we will talk more in the last section of this review. 8. Effects of solar activity at the Earth The Sun, as the central body of our planetary system, has very serious impacts, of various kinds, on interplanetary space and the environments of planets. Near the solar poles the magnetic field lines are open and solar plasma flows continuously into space, creating there the fast solar wind blowing around the Earth deep into outer regions of 7. Accelerated particles Many active processes on the Sun are apparently due to magnetic field-line reconnections, and every reconnection process can accelerate electrons and atomic nuclei to higher energies. In addition to that, eruptions in the chromosphere and corona incite wave motions which propagate both along the solar surface (these waves are often called Moreton waves) and upward through the corona into interplanetary space. Some of these disturbances develop into shock waves, which can be another source of particle acceleration, or can produce second-step acceleration of particles accelerated earlier elsewhere on the Sun. Therefore, the active Sun is a rich source of energetic particles, mainly electrons, protons, and helium nuclei, which in some events can reach energies of hundreds of MeV, and exceptionally particles are recorded even in the BeV range. The most energetic source of accelerated particles are eruptive flares and CME-associated shocks, but also confined flares are sometimes sources of intense flows of accelerated particles, in particular electrons. Particles accelerated in the solar atmosphere are responsible for various kinds of radio emission recorded from the Sun. Clouds of particles, captured in magnetic traps above active regions, cause radio noise storms on metric waves which can last for many days. They have been called Type I radio bursts. Shock waves from flares produce travelling radio disturbances, called Type II bursts. Very frequent radio events are short-lasting Type III bursts which are caused by accelerated electrons streaming along open magnetic field lines high into the corona and in particularly 1508 Figure 11. Above: the development of a coronal mass eject Mission spacecraft. Below: a more recent observation of a co board SOHO on 5 November 1996. The Sun is hidden behind th CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Figure 12. Images of the solar corona in the maximum (above) and minimum (below) of solar cycle. Many coronal helmet streamers dominate the maximum corona. the planetary system. At lower latitudes, coronal helmet streamers (like that one shown in Figure 9 and those in Figure 12) and active regions during periods of field-line openings are sources of slow solar wind. Streams of accelerated particles, both electrons and atomic nuclei, propagate at various places through interplanetary space. And in addition to these streams of plasma and particles, coronal mass ejections send shock waves and plasma clouds in various directions through interplanetary space and eventually cause other particle accelerations there. All this creates highly variable and very complex conditions in the space between the Sun and the Earth and in the last decade we began to speak about, and regularly study, the space weather. In particular, during periods of high solar activity weather in space is very stormy, as one can imagine from a look at Figure 12, which shows the solar corona at the maximum and minimum of the solar activity cycle. For us, of course, most important is the impact at the Earth itself. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 When a flare appears on the Sun, it is the source of X-rays which influences Earth’s ionosphere and thus causes disturbances in radio communications around the Earth. A major eruptive flare can disturb radio contacts for many hours. This disturbance is the same for flares occurring everywhere on the visible solar surface. If a flare (or another active phenomenon on the Sun) accelerates particles, they also can arrive at the Earth, but not from all positions on the solar surface, because they are guided by interplanetary magnetic field lines which – due to the solar rotation – are curved into Archimedean spirals. The most intense particle events come from about 45° west from solar central meridian, and generally from the western solar hemisphere. The most energetic flares emit protons with energy exceeding 500 MeV which arrive at the Earth some 15 min after the flare onset, produce streams of neutrons in Earth’s atmosphere, and cause the so-called ground level effect. Flares that produce protons of such high energies are sometimes called cosmic ray flares. Flares that emit protons with energies higher than 10 MeV are often called proton flares. Particles of lower energy are guided by the Earth magnetic field to the polar regions and cause there absorption of radio waves (polar cap absorption) and intense aurorae. All these effects are delayed by tens of minutes or several hours after the flare onset, depending on the energy of the propagating particles. And then, moving with much slower speeds of a few hundred to 1000 km/s, a coronal mass ejection, often with a shock wave, arrives at the Earth, if it propagates in the right direction. This arrival – two or three days after its origin on the Sun – has a strong impact at the Earth magnetosphere and causes a geomagnetic storm which sometimes can last for several days and has serious impact on communications all around the Earth. Without any doubt, active processes on the Sun also influence the weather at the Earth, but these effects are indirect – depending on the behaviour of the magnetosphere and ionosphere – and very complex: the same effect on the Sun can have quite different consequences at different places of the Earth. Therefore, we still know very little about it. But the active Sun is surely a very important factor in our life. ACKNOWLEDGEMENTS. Illustrations used in this review were obtained at the Big Bear Solar Observatory, California, USA (courtesy Dr H. Zirin); Ottawa River Solar Observatory, Canada (courtesy Dr V. Gaizauskas); University Observatory Wroclaw, Poland (courtesy Prof. B. Rompolt); Skylab Mission (courtesy AS&E, Cambridge, Massachusetts, USA); Yohkoh satellite (courtesy Yohkoh SXT Team), Solar Maximum Mission satellite (courtesy High Altitude Observatory, Boulder, Colorado, USA); and the Large-Angle Spectroscopic Coronagraph (LASCO) on board the SOHO spacecraft (courtesy LASCO consortium). Most illustrations were reprinted from Solar Physics published by Kluwer Academic Publishers in Dordrecht, Holland. 1509 SPECIAL SECTION: SOLAR PHYSICS Probing the Sun’s hot atmosphere Kenneth J. H. Phillips* and Bhola N. Dwivedi**,† *Space Science Department, Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 0QX, UK **Department of Applied Physics, Institute of Technology, Banaras Hindu University, Varanasi 221 005, India The solar corona is an extremely hot (106 K), almost fully ionized plasma which extends from a few thousand km above the photosphere to where it freely expands into the solar system as the solar wind. The exact reason for its high temperature is still unknown, despite more than 50 years of research, but magnetic fields are certainly involved. This article reviews some recent progress in our understanding using data from spacecraft (SOHO, Yohkoh, and TRACE) as well as ground-based eclipse experiments. 1. History of coronal studies Against the expectations of the Second Law of Thermodynamics, the outer atmosphere of the Sun is hotter than the visible surface or photosphere, from which much of its radiation is emitted. The chromosphere, that part of the Sun’s atmosphere which is up to 900 km above the photosphere, has a temperature which rises from a value of 4400 K, the temperature minimum (a region some 500 km above the photosphere), to temperatures of up to 20,000 K, where the dominant radiation is the Lyman lines of neutral hydrogen. Extending above the chromosphere and into the interplanetary space is the solar corona, an extremely hot, tenuous part of the Sun’s atmosphere where the temperature is typically (1 to 2) × 106 K, locally much more. Both the chromosphere and corona are highly structured, with clear evidence of association of magnetic fields which are revealed at the photospheric layers by the Zeeman splitting of magnetically sensitive Fraunhofer lines. The coronal structures are generally loops or large arches, with footpoints apparently in the photosphere, but there are also radial structures called streamers which extend out to very large distances. Over the polar regions, almost radial structures known as plumes are present, giving the whitelight corona during total eclipses the appearance of a bar magnet’s field pattern. This is an important clue to the physical properties of the solar corona. The corona’s high temperature means that it is visible at ultraviolet and X-ray wavelengths, and so many spacecraft and rocket instruments over the years have studied its character. However, it is also visible to the naked eye during the rare circumstances of a total eclipse, when the Moon covers the bright photosphere. The corona then appears as a pearly white, often irregularly shaped, structure all round † the Moon’s limb (Figure 1). The white-light emission is mostly due to Thomson-scattered photospheric light off fast-moving free electrons in the corona. The spectrum of this radiation – the so-called K corona – has the broad characteristics of the photosphere’s spectrum but without the Fraunhofer lines, since they have line profiles which are so highly Doppler-broadened that they cannot be made out against the continuous spectrum. There is a faint extra component, the F corona, due to dust particles in interplanetary space which scatter photospheric radiation also. In this case, the cold dust particles faithfully reproduce the photospheric spectrum including the Fraunhofer lines. The first clues that the corona might be an unusually hot environment were obtained during total eclipses in the nineteenth century. Astronomers were motivated to go to eclipses, even if they were only visible in remote parts of the Earth, as there was no other means available then of studying the Sun’s outer atmosphere. The Americans Young and Harkness studied the corona during the 1869 total eclipse and found a bright emission line at 530.3 nm (to become known as the ‘green’ line) which was unknown in laboratory spectra1. Several more unidentified lines became evident in spectra obtained in subsequent eclipses, and a new element named ‘coronium’ was suspected to be the reason for these spectral lines. As the years passed, however, the periodic table of elements began to be much better understood and it was clear that coronium (as well as ‘nebulium’, discovered from spectral lines in certain gaseous nebulae) could not be easily admitted into the scheme. It was not until the 1930s and 1940s were the lines of coronium and nebulium reproduced in the spectra of very hot spark sources in the laboratory, so it was then realized that the corona must have a temperature of at least that in sparks, nearly a million degrees K. The clinching argument was the discovery by Edlén that the green line was due, not to an unknown element, but to 13-times-ionized iron. Edlén’s work also led to the identification of other coronal lines as being due to multiply ionized atoms of familiar elements like Fe and Ni (see, Phillips1 for more details). 2. The spacecraft era The solar corona is a strong radiator in the ultraviolet and X-ray parts of the spectrum. This radiation is absorbed by For correspondence. (e-mail: dwivedi@banaras.ernet.in) CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1511 SPECIAL SECTION: SOLAR PHYSICS Figure 1 a. Computer-processed image of the total solar eclipse in India on 24 October 1995 (Courtesy E. Hiei). Figure 1 b. Images of part of the west limb of the Sun as follows: (left) From the EIT instrument on the SOHO spacecraft, He II (304 Å) image; (right) From the EIT instrument on the SOHO spacecraft, Fe XII (195 Å) image; (centre) One of 6364 images obtained with the SECIS instrument in Shabla, Bulgaria, during the total solar eclipse of 11 August 1999. Prominences in the SECIS image can be seen in the EIT He II image (emission temp. 20,000 K) while coronal loops above sunspot regions can be seen in the EIT Fe XII image (emission temp. 106 K). 1512 CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS the Earth’s atmosphere so instrumentation has to be flown on rockets or satellites to be able to view it. Observations were begun shortly after World War II by US groups, notably at the Naval Research Laboratory1. Tousey and colleagues obtained the first ultraviolet spectra of the solar atmosphere, finding that the hydrogen Lyman-alpha line, emitted by the chromosphere, was a strong feature at 121.6 nm. Solar X-ray emission was first detected by Burnight in 1949 using a pinhole camera on board a rocket. Spacecraft built by the USA and Soviet Union in the 1960s and 1970s dedicated to solar observations added much to our knowledge of the solar atmosphere, with the manned Skylab Mission of 1973–1974 providing an enormous boost. Ultraviolet and X-ray telescopes on board gave the first high-resolution images of the chromosphere and corona and the intermediate transition region (temperatures between 104 and 106 K). Images of active regions revealed a complex of loops which varied appreciably over their lifetimes, while ultraviolet images of the quiet Sun showed that the transition region and chromosphere followed the ‘network’ character previously known from Ca II K-line images. (The Ca K-line is a visible-wavelength line formed in the chromosphere.) The X-ray images showed that the quiet-Sun corona was characterized by diffuse large-scale loops. In more recent times, the spatial resolution of spacecraft instruments have steadily improved to extremely impressive levels, almost comparable to what can be achieved with ground-based solar telescopes. The Japanese Yohkoh spacecraft, launched in 1991, has on board the US/Japanese soft X-ray telescope (SXT), which images X-rays from active and quiet-Sun regions and flares (which are sudden releases of energy in active regions) with a resolution of about 2 arc seconds (1 arc second corresponds to 725 km at mean solar distance: the mean solar diameter is 32 arc minutes). X-rays with wavelengths in the range 0.2–2 nm are sensed by the SXT. Yohkoh, which is in a low-Earth orbit, continues to operate at the present time, and has obtained many thousands of images from the SXT (Figure 2) as well as considerable amounts of data from the other instruments on board which are mainly for detecting X-ray emission during flares. The ESA/NASA Solar and Heliospheric Observatory (SOHO) was launched in 1995 into an orbit about the inner Lagrangian (L1) point situated some 1.5 × 106 km from the Earth on the sunward side. Its twelve instruments therefore get an uninterrupted view of the Sun, unlike the instruments on Yohkoh. Apart from a period in 1998 when the spacecraft was temporarily out of contact, there have been continuous operations since launch. There are several imaging instruments, sensitive from visible-light wavelengths to the extreme-ultraviolet. The Extreme-ultraviolet Imaging Telescope (EIT), for instance, uses normal-incidence optics to get full-Sun images several times a day in the wavelength bands containing lines emitted by the coronal ions Fe IX/Fe X, Fe XII, Fe XV (emitted in the temperature range 600,000 K to 2,500,000 K) as well as the chromospheric He II 30.4 nm line (Figure 3). The spatial resolution is about 2 arc seconds (1500 km). The set of three coronagraphs making up the Large Angle and Spectrometric Coronagraph (LASCO) view the white-light corona with high resolution out to distances of 30 solar radii (1 solar radius is 700,000 km). Movies of the corona from LASCO show the large-scale structures in the corona as they rotate with the rest of the Sun (the solar rotation period as viewed from the Earth is 27 days or so, with slight latitude dependence), but more particularly they show the large ejections of coronal Figure 2. Yohkoh soft X-ray telescope (SXT) image of the Sun’s corona on 8 May 1992 (Courtesy The Yohkoh SXT Team). CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1513 SPECIAL SECTION: SOLAR PHYSICS mass in the form of huge bubbles, moving out with velocities of several hundred kilometer per second (typically) that, on colliding with the Earth and other planets in the solar system, give the well-known magnetic storms and associated phenomena that have become a matter of widespread concern for telecommunications in recent years. The Transition Region and Coronal Explorer (TRACE) is a highly successful spacecraft observatory, operated by the Stanford-Lockheed Institute for Space Research, which was launched in 1998 and has imaging capabilities that are truly staggering. The spatial resolution is of order 1 arc second (725 km), and there are wavelength bands covering the Fe IX, Fe XII, and Fe XV lines which EIT observes as well as the Lyman-alpha line at 121.6 nm. Movies made by stringing together many images of, for example, active regions reveal a vast wealth of detail, with the coronal loops showing continuous brightenings and motions. A remarkable feature is that the loops often have widths which are no more than the spatial resolution of TRACE, and so they are probably thinner than 1 arc second. Some of the SOHO instruments are able to obtain spectra in the ultraviolet region, and such spectra are of great use in ‘diagnosing’ (i.e. deducing the prevailing physical conditions in) the emitting plasma. The corona is an almost fully ionized plasma, i.e. is composed of mostly protons and electrons, with the density of heavier ions only 10–6 of the proton density (hydrogen is by far the most abundant element in the Sun). However, the atoms of heavier elements like Fe or Si generally retain a few of their electrons, and hence the spectrum of the corona in, e.g. the extreme- 1514 ultraviolet range is characterized by numerous emission lines. The corona is generally optically thin in these lines, and their intensities often give important information about temperatures, densities, and flow speeds. Much work has gone into deducing particle densities from the ratios of particular lines which are density-sensitive, and as a result, we are now able to overlay density maps onto images of portions of the corona. We will discuss this further in Section 5. 3. Heating of the corona: Theory There is evidence that the corona is heated by its magnetic fields, although the evidence is not direct. One vital piece of information that we are still unable to measure is the magnetic field in the corona. We are able to measure, with considerable accuracy, the photospheric magnetic field, using magnetographs that work on the Zeeman principle. This can be done for small regions so that a complete magnetic field map of the Sun’s visible hemisphere (magnetogram) can be constructed. These are routinely available in, for example, the Solar-Geophysical Data Bulletin issued by NOAA. For vector magnetographs, all three components of the magnetic field can be deduced. But the measurements refer to magnetic field at the photospheric level, not in the corona. Although eventually infrared measurements may change this situation, the only way at present in which the coronal field can be deduced is through extrapolations of the photospheric field through the assumption, e.g. of a potential (current-free) or force- Figure 3. Image taken in the Fe XII 19.5 nm extreme-ultraviolet (EUV) line by the SOHO extreme-ultraviolet imaging telescope (EIT) instrument on 7 June 1999 (Courtesy The SOHO EIT Team). CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS free field. We do find, however, that the photospheric field in active regions is more complex than in quiet regions, and it is also known that the active region corona is appreciably hotter (typically 4 × 106 K, depending on the nature of the active region) than in quiet regions (2 × 106 K, less in coronal holes at the poles). So there does seem to be a relation between field strength and heating. A considerable theoretical problem with magnetic field heating is the fact that it requires the diffusion, and therefore reconnection, of magnetic field which implies a resistive plasma. However, the coronal plasma is on the contrary highly conducting. Using Spitzer’s expression for plasma resistivity, η = 103 T – 1.5 Ohm m (T is the temperature in K), we find for T = 2 × 106 K, η = 4 × 10–7 Ohm m, only a factor 20 or so higher than solid copper, a familiar example of an almost perfect conductor, at room temperature! Using the induction equation of magnetohydrodynamics, we find that the diffusion time for a magnetic field is extremely long unless the characteristic distance over which diffusion occurs is as short as a few meters when the diffusion time is a few seconds. Put another way, the magnetic Reynolds number Rm, measuring how tied the magnetic field is to the plasma, is typically 106 to 1012 for the corona, indicating that the field is completely ‘frozen in’ to the plasma. However, reconnection requires Rm to be very small, much less than one in fact. Thus, only if the length scales are very small can one achieve magnetic reconnection. Very small length scales do occur in the region of neutral points or current sheets, where there are steep magnetic field gradients which give rise to large currents. It is thought, then, that such geometries are important for coronal heating if this is by very small energy releases, known (from the original work on the subject, by Parker2) as nanoflares. Some 1016 J are released in a nanoflare, i.e. 10–9 of a large solar flare, and many energy releases like this occurring all over the Sun, quiet regions as well as active regions, could account for heating of the corona. However, it is doubtful whether this mechanism would apply to coronal hole regions where the field lines are open to interplanetary space. The above reasoning applies equally to the competing wave heating hypothesis of coronal heating, in which magnetohydrodynamic waves generated by photospheric motions (e.g. granular or supergranular convection motions) are dampened in the corona. In this case, we need conditions such that the magnetic field changes occur in a shorter time than, say, the Alfvén wave transit time across a closed structure like an active region or quiet Sun loop. The literature for wave heating of the corona is very considerable, but we may briefly summarize it by stating that the waves, generated by turbulent motions in the solar convection zone or at the photosphere, may be surface waves in a loop geometry, or body waves which are guided along the loops and are trapped. The work of Porter, Klimchuk and Sturrock3 shows that short-period fast-mode and slow-mode waves (periods less than 10 s) could be CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 responsible for heating since only for them are the damping rates high enough. It is worth mentioning that MHD waves need not be generated by photospheric motions. Axford and McKenzie4 have proposed a ‘furnace’ model in which small loops convected into the chromospheric network undergo reconnection and launch high frequency waves which can heat the corona (to a few million degrees) through ioncyclotron resonance dissipation and rapidly accelerate the wind up to high speeds (~ 750 km/sec) within a few solar radii. 4. Observational evidence: Transient brightenings Early observations with the high resolution telescope and spectrograph (HRTS) by Brueckner and colleagues5 showed that the profiles of the strong C IV (transition region) ultraviolet (154.8, 155.1 nm) line pair, emitted at 100,000 K, showed much dynamic activity that could be broadly classified into turbulent events (speeds up to 250 km/s in small areas) and jets (speeds up to 400 km/s). The energy contained in the jets amounts to a ‘microflare’ (i.e. up to 1019 J), and it was considered that shock waves generated by a jet could heat up the corona. Enough energy and mass are contained in the jets, assumed to occur over the whole Sun, to satisfy the requirements of the corona and its dynamic extension, the solar wind. Such phenomena are just an example of the many transient events that occur in the solar atmosphere. Shimizu and colleagues6 have been studying the Yohkoh SXT data for active regions, and they find numerous small brightenings in active region loop structures having energies of the order 1020 J, i.e. comparable to microflares. Similar X-ray flares have been noted at higher energies by Lin and colleagues in 1984. Again there is a possibility that the energy supplied by these small active region events is sufficient to heat the corona outside coronal holes, though present indications are that it is short by a factor of about 5. Even smaller events – ‘network flares’ – have been noted outside of active regions by Benz and Krucker and later by Pres and Phillips7 (Figure 4) in studies of Yohkoh SXT and SOHO EIT data. Here, the energies of the events are much less (down to only 100 times a nanoflare) but then the energy requirements of the quiet (i.e. non-active region) solar corona is correspondingly less. Within coronal holes, where the soft X-ray background is very small indeed, Koutchmy has seen tiny coronal ‘flashes’ which have energies of order 10 times a nanoflare. In the extreme-ultraviolet (EUV), very small brightenings have been noted by various authors using SOHO data in quiet-Sun regions. It would appear that these were visible in earlier spacecraft data such as those from the Figure 4. Examples of soft X-ray quiet Sun ‘network flares’, noted o 1515 SPECIAL SECTION: SOLAR PHYSICS OSO series from the 1970s. A comparison by Pres of network flares and these so-called ‘blinkers’ reveals a strange lack of correlation, with the EUV transients in quiet regions being apparently much more numerous than their Xray counterparts. Table 1 gives some indication of the thermal energy rate (W) deduced from observations of various transients in either X-rays or the EUV. As the total radiative power of the entire quiet-Sun corona is several 1019 W, it appears that the numerous EUV quiet-Sun transients might offer the best possibility of heating, assuming that the Parker nanoflare hypothesis is correct. 5. Physical characteristics of the corona It is clear from eclipse or spacecraft images of the corona (Figures 1 and 2) that the corona is highly structured, and that the hot plasma making up the corona is confined to intricate magnetic field patterns. Particle densities and temperatures can be derived for different coronal regions using a variety of methods. For example, measurements of the surface brightness of the white-light corona yield electron densities, since the emission is by Thomson scattering of photospheric light off free electrons. Near the base of the corona the measured electron densities are a few times 1014 m–3, though this is strongly dependent on whether the feature is a quiet region (smaller densities) or within a complex of active region loops (larger densities). The densities are further reduced within coronal holes. The density falls off rapidly with height: at 5 solar radii from Sun centre, the density is 1011 m–3, while at the distance of the Earth’s orbit (where the corona is in the form of a freely flowing wind) the density is less than 107 m–3. As indicated earlier, temperatures vary in the corona from place to place, with maximum values in highly complex active regions (up to 4 × 106 K) and minimum values in the polar coronal holes (slightly less than 106 K). In general, then, densities and temperatures are correlated. To illustrate the fact that the coronal gas is strongly dominated by the magnetic field, we find that the magnetic pressure B2/2 µ is generally many times more than the gas pressure NkT (N = particle number density, k = Boltzmann’s constant). For a typical active region, magnetic pressure might be as much as 50 Pa, but the gas pressure is perhaps Table 1. Energetics for EUV and X-ray transient events in the quiet Sun corona No. of events h –1 40,000 brightenings 1,200 100 Thermal energy rate (W) 3 × 10 2 × 10 18 2 × 10 18 19 Event type EUV transients Small network flares Large network flares The data in the table are based on recent publications using SOHO and Yohkoh observations (see References). 1516 a factor of almost 10 less. Measurements of densities and temperatures using X-ray or ultraviolet data are possible, particularly using line intensities. Most regions in these lines are optically thin, though there are notable exceptions like in the hydrogen Lyman-alpha line and the other strong resonance lines. The line intensity is a function of temperature, through the ionization fraction of the emitting ion (a strongly peaked curve) and the excitation rate from the ground energy level of the ion to the upper level giving rise to the line emission. Excitation of most ions emitting lines in the X-ray and ultraviolet ranges is due to collisions of free electrons with the ions. Very roughly we can assign a single temperature to an ion, e.g. three-times-ionized carbon, which gives rise to the C IV 154.8/155.1 nm line doublet, corresponds to a temperature of about 100,000 K, approximately the maximum fractional abundance of the C+3 ion in ‘ionization equilibrium’, i.e. a balance between collisional ionization processes and radiative and dielectronic recombination processes which occurs in the solar corona. These fractional abundances can be calculated using atomic data, and there are many publications which give values as a function of temperature (in general there is practically no density dependence). Excitation of the lines can be calculated to much higher accuracy than was possible about 20 years ago as there are sophisticated atomic codes which take into account resonances in the collisional cross sections. Among these is the close-coupling R-matrix code developed at Queen’s University Belfast. As a result, there are a number of line pairs recognized in especially the extreme-ultraviolet part of the spectrum which are sensitive to electron densities. This fact is very useful as nearly all other methods of getting densities are indirect. Thus, a measured line intensity of N e2aV feature leads to a value for the ‘emission measure’ (V = volume, Ne = electron number density). This combined with a measured value for V (e.g. from image data) gives Ne . However, this technique is often quite imprecise since the presence of fine structure within the feature, if unresolved by the instrument taking the observations, renders the value of Ne to be merely a lower limit. Account then has to be taken of a ‘filling factor’, often much less than one. Spectral line diagnostics for solar plasmas have been discussed in an accompanying article of this issue by Dwivedi, Mohan and Wilhelm8. The Coronal Diagnostic Spectrometer (CDS) on SOHO has been widely used to get densities, and it is now possible to construct maps of regions of the Sun showing electron densities. Gallagher and colleagues9 at Queen’s University Belfast and Rutherford Appleton Laboratory have been active in this. Two examples of suitable densitysensitive lines in the wavelength range of CDS are those due to Si IX and Si X, both emitted at around 1.3 × 106 K. This value of temperature makes them ideal for studying the quiet corona. Figure 5 shows the positions of scans using a pair of lines emitted by each ion around the Sun’s limb with CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS the CDS instrument in February 1996, and Figure 6 shows the resulting scan with light areas indicating measured electron densities from each ion, the vertical scale being position angle around the Sun’s limb. It shows clearly the presence of higher densities in low-latitude regions (Ne around 4 × 1014 m–3), while near the south pole (position angle 180 degrees) the electron density is at least a factor 4 lower. The density profile with position angle at three radial distances out to 1.2 solar radii agrees remarkably well with a recent analytical model that has been developed for Sun-like stars 10. 6. Observational evidence: Wave motions Despite the fact that the nanoflare hypothesis seems to be observationally plausible (as indicated in Section 4), MHD waves may well be implicated in the heating of the corona, and it is important to look for signatures of them. It is, for example, unlikely that nanoflares could heat the corona in the regions of open field lines such as occur in polar coronal holes, yet it appears that the corona is still hot in these open-field regions. A basic difficulty seems to be that theoretical predictions indicate that MHD waves having only a short period (less than 10 sec) are likely to be effective in the heating, since only for such waves are the damping rates sufficiently great. However, spacecraft imaging, limited as it is by the rate at which data are telemetered to the Earth, is necessarily rather slow. It takes about 2 min for any instrument on SOHO to produce an image of even relatively small portions of the Sun. There is hence still considerable interest in observing the visible-light corona during total eclipses from the ground, since one can use high-speed electronic cameras to obtain rapid imaging of particular coronal structures. A pioneer in this work has been Pasachoff 11, who has tried this kind of experiment at various eclipses around the world since the 1980s. Analysis of his best results indicate the presence of a slight peak in Fourier spectra at frequencies of 0.5–1 Hz. This has been seen in more recent eclipses, including the 1998 eclipse in the Caribbean. Other measurements using ground-based white-light coronagraphs have been taken, notably by Koutchmy12 at the US National Solar Observatory/Sacramento Peak some years ago. Here searches were made for periodic modulations in both the intensity and velocity of the green line, with evidence of periods equal to 43, 80 and 300 s (the last is probably related to the familiar five-minute oscillation seen with photospheric Fraunhofer lines). While Pasachoff continues to develop his instrument with colleagues at Williams College, Massachusetts, a group including Rutherford Appleton Laboratory, Queen’s University Belfast, and the Astronomical Institute, University of Wroclaw in Poland have been developing a fastimaging system with charge-couple device (CCD) cameras that can image up to 50 frames a second with a specially adapted computer that ‘grabs’ images, placing the data on to large-capacity hard discs for later analysis. The cameras were developed by EEV, a company specializing in CCD cameras in Chelmsford, UK, and the computer hardware and software were developed by Carr-Crouch Computer Company, Maidenhead, UK. The system, called the solar eclipse coronal imaging system (SECIS), has been tested on a number of occasions already, the first being during the 1998 eclipse and most recently on 11 August 1999, the last total solar eclipse of this millennium. Scientifically useful results were obtained during a run on the Evans Coronagraph Facility at Sacramento Peak. Figure 6. Electron density plot showing variation of electron density with position angle round the Sun (see Figure 5). Density Figure 5. SOHO Figure EIT7.image Image of of thecoronal Sun showing loops on positions the solar of scans limb in September scale is on1998 the right. with The preliminary left panelFourier is using analysis a Si IXof line tworatio, points thatinon the image (A is with the CDS instrument. within the coronal active region, B is considered to be non-coronal the rightorasky Si Xbrightness) ratio (Courtesy (Courtesy P. T. P. Gallagher). T. Gallagher). CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1517 SPECIAL SECTION: SOLAR PHYSICS Preliminary results, with one channel (using a green-line filter), are shown in Figure 7, showing a system of active region loops on the limb with initial Fourier analysis done by Gallagher. Taken at face value, there is very slight evidence for excess power at small (less than 5 Hz) at a location (A) within the active region loops but none in the more distant location (B) where the coronal signal is negligible compared with sky (terrestrial) background. 7. Concluding remarks A large amount of information concerning the physics of the solar atmosphere has become available in recent years and it is taking some time to digest the new data. The exact reason for heating the solar corona is still not known for certain, but there is much evidence from spacecraft observations that, in regions where there are complex magnetic field geometries or at least the presence of closed loops, heating by numerous small flare-like releases of energy is adequate to explain the energy requirements of the corona. In coronal holes, i.e. regions of open field lines, this is less likely to be true and many consider that heating proceeds through the damping of MHD waves, which may still have a role in the heating of the corona in closed-field regions. At present this can only be investigated using ground-based instruments since the periods of MHD waves which have sufficiently large damping rates are likely to be very small, of order a few seconds. Spacecraft imaging is too slow to search for such periodicities. 1518 1. Phillips, K. J. H., Guide to the Sun, Cambridge University Press, Chapter 5, 1995. 2. Parker, E. N., Astrophys. J., 1988, 330, 474–479. 3. Porter, L. J., Klimchuk, J. A. and Sturrock, P. A., Astrophys. J., 1994, 435, 482–501. 4. Axford, W. I. and McKenzie, J. F., in Cosmic Winds and the Heliosphere (eds Jokipii, J. R., Sonett, C. P. and Giampapa, M. S.), University of Arizona Press, Tucson, 1997, pp. 31–66. 5. Brueckner, G. E. and Bartoe, J.-D. F., Astrophys. J., 1983, 272, 329–348. 6. Shimizu, T., Pub. Astron. Soc. Japan, 1995, 47, 251–263. 7. Pres, P. and Phillips, K. J. H., Astrophys. J., 1999, 510, L73– L76. 8. Dwivedi, B. N., Mohan, A. and Wilhelm, K., Curr. Sci., 1999 (this issue). 9. Gallagher, P. T., Mathioudakis, M., Keenan, F. P., Phillips, K. J. H. and Tsinganos, K., Astrophys. J. (submitted). 10. Lima, J. J. G., Priest, E. R. and Tsinganos, K., The Corona and Solar Wind near Minimum Activity, ESA SP-404, 1997, pp. 521–526. 11. Pasachoff, J. and Ladd, E. F., Solar Phys., 1987, 109, 365–372. 12. Koutchmy, S., Zugzda, Y. D. and Locans, V., Astron. Astrophys., 1983, 120, 185–191. ACKNOWLEDGEMENTS. We thank colleagues for allowing us to use their work to illustrate this article, in particular P. T. Gallagher, F. P. Keenan, P. Pres, and members of the SOHO (EIT) and Yohkoh experiment teams. We are indebted to K. Wilhelm for critical reading of the manuscript. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS A new Sun: Probing solar plasmas in the extreme-ultraviolet light from SUMER on SOHO Bhola N. Dwivedi*,**,†, Anita Mohan* and Klaus Wilhelm** *Department of Applied Physics, Institute of Technology, Banaras Hindu University, Varanasi 221 005, India **Max-Planck-Institut für Aeronomie, 37191 Katlenburg-Lindau, Germany We briefly outline the extreme-ultraviolet spectroscopy of solar plasmas in the light of a wealth of high-resolution observations, both in spectral and spatial regimes, from the SUMER spectrograph (Solar Ultraviolet Measurements of Emitted Radiation) on the spacecraft SOHO (Solar and Heliospheric Observatory). We then present some of the new results on a new Sun seen by the SUMER spectrograph. In particular, we discuss coronal holes and the solar wind, the ‘red’/‘blue’ Sun, abundance anomalies, explosive events and sunspot transition region oscillations. We conclude this article by reiterating clues, obtained from SUMER, that provide information essential for solving solar riddles of coronal heating and the wind acceleration. 44 mÅ in first order of diffraction and 22 mÅ in second order. The spectral resolution, which can be improved for relative (and under certain conditions for absolute) measurements to fractions of a pixel, allows to measure Doppler shifts corresponding to plasma bulk velocities of about 1 km s –1 along the line of sight. Turbulent velocities are obtained by determining line broadenings. Vast literature has resulted on the basis of observations carried out using SUMER. But, we have not reviewed this in the present article. Instead, we present some new results from SUMER on solar plasmas that have added a new dimension to a better understanding of some of the solar mysteries, especially coronal heating and the solar wind acceleration. Introduction Plasma diagnostics THE Sun presents us with a thousand-fold face. Depending on the ways we observe it, whether it be through a groundbased telescope, during an eclipse, or from a space observatory, we see a different Sun. One can distinguish the wavelengths in which it is observed: X-rays, ultraviolet, visible, infrared, radio or even by non-photon instruments (e.g. neutrino detectors). One can distinguish the time when it is observed: near the maximum activity of its sunspot cycle, or the minimum activity. Still this multi-faceted Sun of ours is a single celestial object, and it is this idea of the Sun as whole that has been described in the previous articles of this special issue. A new Sun that has emerged from the analysis and the interpretation of observations from the Solar Ultraviolet Measurements of Emitted Radiation (SUMER) spectrograph is briefly presented in this article. Full descriptions of the SUMER spectrograph on the spacecraft SOHO and its performance are available1–3. Briefly, the SUMER spectrograph observes the Sun in the extreme-ultraviolet (EUV) light from 465 Å to 1610 Å with high spatial and spectral resolution. This wavelength range contains EUV lines from the chromosphere, the transition region, and the corona, thereby providing a unique opportunity to probe plasmas of the solar atmosphere. The spatial resolution is close to 1″ (about 715 km at the Sun), while the spectral resolution element (one pixel) is about Without a knowledge of the densities, temperatures, and elemental abundances of space plasmas, almost nothing can be said regarding the generation and transport of mass, momentum and energy. Thus, since early in the era of space-borne spectroscopy we have faced the task of inferring plasma temperatures, densities, and elemental abundances for hot solar and other astrophysical plasmas from optically thin emission-line spectra4,5. A fundamental property of hot solar plasmas is their inhomogeneity. The emergent intensities of spectral lines from optically thin plasmas are determined by integrals along the line of sight (LOS) through the plasma. Spectroscopic diagnostics of the temperature and density structures of hot optically thin plasmas using emission-line intensities is usually described in two ways. The simplest approach, the line-ratio diagnostics, uses an observed line-intensity ratio to determine density or temperature from theoretical density or temperature-sensitive line-ratio curves, based on an atomic model and taking account of physical processes for the formation of lines. The line-ratio method is stable, leading to well-defined values of Te or Ne , but in realistic cases of inhomogeneous plasmas these are hard to interpret, since each line pair yields a different value of density or temperature. The more general differential emission measure (DEM) method recognizes that observed plasmas are better described by distributions of temperature or density along the LOS, and poses the problem in inverse † For correspondence. (e-mail: dwivedi@banaras.ernet.in) CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1521 SPECIAL SECTION: SOLAR PHYSICS form. It is well known that the DEM function is the solution to the inverse problem, which is function of Te , Ne , or both. Derivation of DEM functions, while more generally acceptable, is unstable to noise and errors in spectral and atomic data. The exact relationship between the two approaches has never been explored in depth, although particular situations were discussed6. The mathematical relationship between these two approaches has recently been reported7. Line shifts and broadenings give information about the dynamic nature of the solar and stellar atmospheres. The transition region spectra from the solar atmosphere are characterized by broadened line profiles. The nature of this excess broadening puts constraints on possible heating processes. Systematic redshifts in transition region lines have been observed in both solar spectra and stellar spectra of late-type stars. On the Sun, outflows of coronal material have been correlated with coronal holes. The excess broadening of coronal lines above the limb provides information on wave propagation in the solar wind. Figure 1 shows the full Sun raster image obtained by SUMER in the Ne VIII (λ770) line on 2 February 1996. The emission line is formed at 630000 K and observed in second order. The polar coronal holes are clearly seen in this line as well as some bright points and polar plumes8. In Figure 2 a Doppler Figure 1. Full Sun raster image obtained by Solar Ultraviolet Measurements of Emitted Radiation (SUMER) in the line Ne VIII (λ770) on 2 February 1996. The emission line is formed at a temperature of 630000 K and observed in second order of diffraction on the bare microchannel plate portion of detector A. The 1″ x 300″ slit was used with a step width of 1.88″ and an exposure time of 7.5 s. The polar coronal holes can clearly be seen in this line as well as some bright points and polar plumes (from Wilhelm et al.8 ). 1522 velocity map is shown, derived from the Ne VIII observations under the assumption that the line shifts can be interpreted as plasma flows. The range of the velocity scale is from – 30 km s –1 (blue) to + 30 km s –1 (red). The zero point is adjusted to give no Doppler shift just above the limb at about 20 arc sec. The blue polar coronal holes stand out with some white spots, which upon close inspection, coincide with bright points in the intensity image. Strong up and down motions can be seen near the active regions in the western hemisphere. Coronal holes and the solar wind In 1950s Waldmeier9 first recognized persistent depressions in the intensity of the monochromatic corona (outside the polar caps) as observed by ground-based coronagraphs and called them ‘holes’ (Löcher in German). He published his coronal observations obtained between 1939 and 1952 (during the solar cycles 17 and 18). At the end of cycle 17, coronal holes were identified in coronal maps. More than 20 years later, after theoretical predictions, Krieger et al.10 related a coronal hole, seen on an X-ray image taken on 24 November 1970 during a sounding rocket flight, to a recurrent high-speed stream of Figure 2. Doppler velocity map derived from the Ne VIII observations shown in Figure 1. The range of the velocity scale from – 30 km s–1 (blue) to + 30 km s–1 (red). The zero point is adjusted to give no Doppler shift just above the limb (at ~ 20″). The blue polar coronal holes stand out with some white spots, which upon close inspection, coincide with bright points in the intensity image. Strong up and down motions can be seen near the active regions in the western hemisphere. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS the solar wind observed by instruments on PIONEER-VI and on the Vela satellite. Observations from the Skylab mission further established that the high-speed solar wind originates in coronal holes which are well-defined regions of strongly reduced EUV and soft X-ray emissions11. More recent data from Ulysses show the importance of the polar coronal holes, particularly at times near the solar minimum when a magnetic dipole dominates the field configuration. Figure 3 shows a polar coronal hole with bright points and polar plumes12 seen in Mg X (λ625) with a formation temperature of ~ 1100000 K. The mechanism for accelerating the wind to the high values observed, of the order of 800 km s –1, is not understood quantitatively. The Parker model13 is based upon a thermally driven wind. To reach such high velocities, temperatures of the order 3 to 4 MK would be required near the base of the corona. However, other processes are available for acceleration of the wind, for example the direct transfer of momentum from magneto-hydrodynamic (MHD) waves, with or without dissipation. This process results from the decrease of momentum of the waves as they enter less dense regions, coupled with the need to conserve momentum of a system consisting of waves plus the local plasma. If this transfer predominates, it may not be necessary to invoke high coronal temperatures at the base of the corona. In reality, very little information was available on the density and temperature structure in coronal holes prior to the SOHO Mission. Data from Skylab is limited, due to the very low intensities in holes and the poor spectral resolution, leading to many line blends. Skylab was able to follow temperatures up to nearly 1 MK and no further, and the interpretation of the data is quite uncertain. Highresolution EUV observations from instruments on SOHO provided the opportunity to infer the density and temperature profile in coronal holes. Wilhelm et al.14 observed several polar coronal holes with SUMER spectrograph in Si VIII lines at 1445.75 Å and 1440.49 Å and Mg IX lines at 706 Å and 750 Å to determine density and temperature structure via line-ratio spectroscopic diagnostics. Figure 4 shows the electron densities deduced. Comparing the electron temperatures with the ion temperatures, Wilhelm et al. concluded that ions are extremely hot and the electrons are relatively cool. This result is also in agreement with the UVCS (Ultraviolet Coronagraph Spectrometer) SOHO results15 at greater altitudes. Using the two SOHO instruments CDS (Coronal Diagnostic Spectrometer) and SUMER David et al.16 have now measured electron temperatures as a function of height above the limb in a polar coronal hole. Observations of two lines from the same ion stage O VI 1032 Å from SUMER and O VI 173 Å from CDS/SOHO were made to determine the electron temperature gradient in a coronal hole. They deduced temperatures of around 0.8 MK close to the limb, rising to a maximum of less than 1 MK at 1.15 R¤, then falling to around 0.4 MK at 1.3 R¤. It seems that present observations preclude the existence of temperatures over 1 MK at any height near the centre of a coronal hole. Wind acceleration by temperature effects is therefore inadequate as an explanation of the high-speed wind and it becomes essential to look towards other effects, probably involving the momentum and the energy of Alfvén waves. The ‘red’/‘blue’ Sun In the so-called transition region, the temperature rises sharply from some 10000 K to more than 1 MK. It has been known since the Skylab period that there is a net red shift in the transition region lines17. It has recently been reported that the redshift peaks at about 12 km s –1 near 150000 K and extends into the hotter regions where Ne VIII emission line is formed18,19. Ne VIII belongs to the Figure 3. Polar coronal hole with bright points and polar plumes seen in Mg X (λ625) with a formation temperature of ~ 1100000 K (from Dammasch et al.12 ). CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1523 SPECIAL SECTION: SOLAR PHYSICS Li-sequence and has a strong 2s–2p resonance line at 770 Å which is observed by SUMER in both first- and secondorder. A laboratory wavelength of (770.409 ± 0.005) Å was used for this line. Dammasch et al.20 made an accurate wavelength measurement of this line recorded in second order together with several S I and C I lines, which have well-known wavelengths. Assuming that there is no net Doppler flow along the line of sight at and above the limb and eliminating other residual errors, a rest wavelength of (770.428 ± 0.003) Å was derived. With 1 mÅ corresponding to 390 m s –1, this new result moves the Doppler shift of Ne VIII towards the blue by (–)7.4 km s –1. This immediately implies that there is no net downflow at this temperature in quiet Sun regions. That the solar wind is coming from coronal holes (open magnetic field regions in the corona) has been widely accepted, although little additional direct observational evidence has been obtained to support this view. Hassler et al.21 found the Ne VIII emission blueshifted in the north polar coronal hole along the magnetic network boundary and at network boundary interfaces compared to the average quiet Sun flow. We show in Figure 5 velocity map of the Ne VIII (λ770) line as observed on 21 September 1996 (top panel). The areas of dark blue, corresponding to an outflow velocity of more than 5 km s –1 (LOS), are enriched by contours. A repeat of the velocity contours is overlaid on the intensity diagram of the same area in the bottom panel. It is to be noted that coronal hole conditions prevail in the northern portion of the field of view (520″ × 300″), whereas quiet Sun regions are present in the south. These Ne VIII observations reveal the first two-dimensional coronal images showing velocity structure in a coronal hole, and provide strong evidence that coronal holes are indeed the source of the fast solar wind. The apparent relationship to the chromospheric magnetic network, as well as the relatively large outflow velocity signatures at the intersections of network boundaries at midlatitudes, is a first step in better understanding the complex structure and dynamics at the base of the corona and the source region of the solar wind. combination/forbidden lines, which provided good possibilities to study the relative element abundance of Ne (high FIP) and Mg (low FIP) in transition region emission in the corona. The observation of the FIP effect in transition region emission in the corona is a new observational fact. Laming et al.26 have also investigated the behaviour of the FIP effect with height above the solar limb in a region of diffuse quiet corona and found a low FIP bias of a factor of 3 to 4, with no significant height variation. Yet another study confirmed the new observational facts about FIP effects in transition region emission27 which is shown in Figure 6 for the Mg/Ne FIP bias from four line ratios as a function of height above the limb. Explosive events The universe abounds with explosive energy release that may heat plasma to millions of degrees and accelerate particles to relativistic velocities. Such occurrences are not Abundance anomalies The status of coronal abundances relative to hydrogen is not entirely settled. It is often suggested that abundances are correlated with the first ionization potential (FIP). Many studies show that ‘FIP bias’ does exist22–24. Classically, a step function increase by a factor of 4 is assumed for elements with increasing FIP. The FIP effect should eventually offer valuable clues to the process of heating, ionization, and injection of material into coronal and flaring loops for the Sun and other stars. Dwivedi et al.25 presented results from a study of EUV off-limb spectra obtained on 20 June 1996 with SUMER spectrograph. They recorded Ne VI and Mg VI inter1524 Figure 5. Velocity map of the Ne VIII (λ770) line as observed on 21 September 1996 (top panel). The areas of dark blue, corresponding to an outflow velocity of more than 5 km s–1 LOS (line-of-sight), are enriched by contours. Single pixel contributions have been eliminated by a floating average over 3″ in east–west and north–south directions. A repeat of the velocity contours is overlaid on the intensity diagram of the same area in the bottom panel. Note that coronal hole conditions prevail in the northern portion of the field of view (520″ × 300″), whereas quiet Sun regions are present in the south. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS uncommon on our own star too. Examples include solar flares, coronal mass ejections, chromospheric and coronal microflares, etc. In many cases, the magnetic field seems to be the only source of energy available to power these cosmic explosions. While it is well established that the Sun has a large reservoir of magnetic energy, the reason for its release is still debated. The widely accepted explanation for explosive energy release is a process known as magnetic reconnection. This effectively involves the cutting and reattachment of magnetic lines of force. Many solar physicists feel that reconnection is observationally and theoretically well established. Sceptics, however, argue that there is no definite proof for the reconnection taking place on the Sun. It is for this reason that new results from SUMER for the magnetic reconnection on the Sun is so important. They provide the best evidence to date for the existence of bidirectional outflow jets, forming fundamental part of the standard reconnection model. Explosive events28 were first seen in the ultraviolet spectra obtained with the NRL’s (Naval Research Laboratory) High Resolution Telescope and Spectrograph Figure 6. (HRTS) flown on several rocket flights and Spacelab 2. They were found to be short-lived (60 sec), small-scale (1500 km), high-velocity (± 150 km s –1) flows that occurred very frequently over the entire surface of the Sun. There are estimated to be 30,000 events at any one time on the Sun. The energy involved in the events observed, however, suggests that they are not the major source of mass or energy in either chromosphere or corona. Their importance lies in the fact that they probably represent the high energy tail of a spectrum of network events that occur on scales unobservable with current techniques. It is also noted that explosive events are associated with freshly emerged magnetic field and their Doppler velocities are roughly equal to the Alfvén speed in the chromosphere. The suggestion is that the events result from magnetic reconnection. Possible evidence for bi-directional nature of the flows had been noted in earlier spectroscopic data but the examples were not very clear. This was so because the structure of the flow could not be resolved due to limited time and space coverage by previous space experiments. The SUMER spectrograph has now made it possible to observe the chromospheric network continuously over an Mg/Ne FIP bias derived from four line ratios as a function of height above the limb (from Dwivedi et al.27 ). CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 1525 SPECIAL SECTION: SOLAR PHYSICS extended period and to discern the spatial structure of the flows associated with these explosive events. The observation that explosive events are bi-directional jets, provides new evidence that they result from magnetic reconnection above the solar surface. From simultaneous magnetic field and ultraviolet measurements, it has already been suggested that explosive events are often found on the chromospheric network boundary and seem to be associated with the cancellation of photospheric magnetic fields. The network consists of curtains of very strong magnetic flux tubes. All flux tubes are anchored by their footpoints to the photosphere. The continual motions in the photosphere mean that field lines of opposite-polarity are naturally drawn together. If flux tubes with oppositepolarity field lines are pushed together, a current sheet forms. In a finite resistivity plasma, a small region near the neutral region may collapse and create a thin reconnection region. From this region, plasma is ejected in both directions along the field lines, with velocity of the order of the Alfvén speed (the Alfvén speed depends on magnetic field strength and plasma density). Electrical resistance to this current flow liberates energy from the system to heat the plasma, much in the same way as the filament of a light bulb is heated. Interpreting the evolution of the jets in the Si IV 1393 Å line profile, Innes et al.29 have now shown that explosive events have the bi-directional jets ejected from small sites above the solar surface. In Figure 7, we show the explosive events seen in the Si IV (λ1393) line. The SUMER spectrograph slit shown in each section has a projected length of 84000 km on the Sun and a width of 700 km. The exposure time was 10 s each. The Doppler shift near bright portions correspond to plasma motions with line-of-sight velocities ± 150 km s –1. The structure of these plasma jets evolves in the manner predicted by theoretical models of magnetic reconnection. This lends support to the view that magnetic reconnection is one of the fundamental processes for accelerating plasma on the Sun. Such observations seem to provide the best evidence to date for the existence of bidirectional outflow jets, a fundamental part of the standard Figure 7. Explosive events seen in the Si IV (λ1393) line. The SUMER spectrograph slit shown in each section has a projected length of 84000 km on the Sun and a width of 700 km. The exposure time was 10 s each. The Doppler shift near bright portions (i.e. at chromospheric network crossings of the slit) correspond to plasma motions with line-of-sight velocities ± 150 km s–1 . 1526 CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 SPECIAL SECTION: SOLAR PHYSICS Sunspot transition region oscillations The sunspot transition region between the chromosphere and the corona oscillates. This may reveal crucial information about the structure and the physics of sunspots. The first detailed study of the oscillations in the sunspot transition region was presented by Gurman et al.30 mainly based on observations of eight sunspots in the C IV (λ1548) line with the UVSP (Ultraviolet Spectrometer and Polarimeter) instrument on the Solar Maximum Mission. They observed oscillations with periods of 129–173 s, with no signs of shocks and suggested that the oscillations are caused by upward-propagating acoustic waves. Observations with instruments on SOHO have revived the interest for the transition region oscillations. Intensity oscillations in the 2 min range were recently reported31. Based on observations with a 15 sec time resolution Fludra32 has observed 3 min-intensity oscillations and suggested that oscillations occur in sunspot plumes. Combining observations of intensity and line-of-sight velocity oscillations in three transition region lines, Brynildsen et al.33 found that their observations of NOAA 8156 were compatible with the hypothesis that the 3 min oscillations in sunspot transition region are caused by linear, upward-propagating, progressive acoustical waves. This appears to be in conflict with the upward propagating shock waves observed in the sunspot chromosphere34. Hence, either the wave amplitude decreases abruptly between the chromosphere and the transition region or considerable difference in the 3 min umbral oscillations exist between different sunspots. Figure 8 shows sunspot oscillations observed with SUMER in the O V (λ629) line. In the upper panel, the position of the spectrometer slit analysed is shown with respect to the sunspot of a TRACE image of NOAA 8487 on 18 March 1999. In the lower panels the intensity variations and the line-of-sight velocity along the slit section are displayed. Figure 8. Sunspot oscillations observed with SUMER in the O V (λ629) line. In the upper panel the position of the section of the spectrometer slit analysed is shown with respect to the sunspot of a TRACE image of NOAA 8487 on 18 March 1999. In the lower panels the intensity variations and the line-of-sight velocity along the slit section are displayed (courtesy, N. Brynildsen and the TRACE team). magnetic reconnection model. There exists a vast collection of data obtained from the SUMER instrument on the spacecraft SOHO and similar bi-directional jets are expected to be seen in the solar atmosphere wherever reconnection takes place. The present and future observations of this kind are likely to provide new clues to a better understanding of how the Sun’s magnetic energy feeds its million-degree hot corona and the solar wind. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999 Concluding remarks In conclusion, SUMER results on plasma density and temperature structure, and evidence for hot ions and cool electrons in coronal holes, the ‘blue’ Sun, observational evidence for magnetic reconnection on the Sun, FIP effect in the corona and sunspot transition region oscillations presented in this article are crucial to a better understanding of how the Sun’s magnetic energy feeds its million-degree hot corona and the solar wind. And it will take some time to digest SUMER data to decipher what tricks the Sun is performing! 1. Wilhelm, K., Curdt, W., Marsch, E. et al., Solar Phys., 1995, 162, 189–231. 1527 SPECIAL SECTION: SOLAR PHYSICS 2. Wilhelm, K., Lemaire, P., Curdt, W. et al., Solar Phys., 1997, 170, 75–104. 3. Lemaire, P., Wilhelm, K., Curdt, W. et al., Solar Phys., 1997, 170, 105–122. 4. Dwivedi, B. N., Space Sci. Rev., 1994, 65, 289–316. 5. Mason, H. E. and Monsignori-Fossi, B. C., Astron. Astrophys. Rev., 1994, 6, 123–179. 6. Brown, J. C., Dwivedi, B. N., Almleaky, Y. M., Sweet, P. A., Astron. Astrophys., 1991, 249, 277–283. 7. McIntosh, S. W., Brown, J. C., Judge, P. G., Astron. Astrophys., 1998, 333, 333–337. 8. Wilhelm, K., Lemaire, P., Dammasch, I. E., Hollandt, J., Schühle, U., Curdt, W., Kucera, T., Hassler, D. M. and Huber, M. C. E., Astron. Astrophys., 1998, 334, 685–702. 9. Waldmeier, M., Die Sonnenkorona I (1951), Die Sonnenkorona II (1957), Birkhäuser Verlag, Basel, Stuttgart. 10. Krieger, A. S., Timothy, A. F. and Roelof, E. C., Solar Phys., 1973, 29, 505–525. 11. Zirker, J. B. (ed.), Coronal Holes and High Speed Wind Streams, Colorado Associated Univ. Press, Boulder, 1977. 12. Dammasch, I. E., Hassler, D. M., Wilhelm, K. and Curdt, W., in 8th SOHO Workshop (ed. Kaldeich-Schürmann, B.), ESA, SP446, 1999, in press. 13. Parker, E. N., Astrophys. J., 1958, 128, 664–676. 14. Wilhelm, K., Marsch, E., Dwivedi, B. N., Hassler, D. M., Lemaire, P., Gabriel, A. H. and Huber, M. C. E., Astrophys. J., 1998, 500, 1023–1038. 15. Kohl, J. L., Noci, G., Antonucci, E. et al., Solar Phys., 1997, 175, 613–644. 16. David, C., Gabriel, A. H., Bely-Dubau, F., Fludra, A., Lemaire, P. and Wilhelm, K., Astron. Astrophys., 1998, 336, L90–L94. 17. Doschek, G. A., Feldman, U. and Bohlin, J. D., Astrophys. J., 1976, 205, L177–L180. 18. Brekke, P., Hassler, D. M. and Wilhelm, K., Solar Phys., 1997, 175, 349–374. 19. Chae, J., Yun, H. S. and Poland, A. I., Astrophys. J. Suppl., 1998, 114, 151–164. 20. Dammasch, I. E., Wilhelm, K., Curdt, W. and Hassler, D. M., 1528 Astron. Astrophys., 1999, 346, 285–294. 21. Hassler, D. M., Dammasch, I. E., Lemaire, P., Brekke, P., Curdt, W., Mason, H. E., Vial, J.-C. and Wilhelm, K., Science, 1998, 283, 810–813. 22. Widing, K. G. and Feldman, U., Astrophys. J., 1993, 416, 392– 397. 23. Sheeley, N. R., Astrophys. J., 1996, 469, 423–428. 24. Young, P. R. and Mason, H. E., Solar Phys., 1997, 175, 523– 539. 25. Dwivedi, B. N., Curdt, W. and Wilhelm, K., Astrophys. J., 1999, 517, 516–525. 26. Laming, J. M., Feldman, U., Drake, J. J. and Lemaire, P., Astrophys. J., 1999, 518, 926–936. 27. Dwivedi, B. N., Curdt, W. and Wilhelm, K., in 8th SOHO Workshop (ed. Kaldeich-Schürmann, B.), ESA SP-446, 1999 (in press). 28. Dere, K. P., Bartoe, J.-D. F. and Brueckner, G. E., J. Geophys. Res., 1991, 96, 9399–9407. 29. Innes, D. E., Inhester, B., Axford, W. I. and Wilhelm, K., Nature, 1997, 386, 811–813. 30. Gurman, J. B., Leibacher, J. W., Shine, R. A., Woodgate, B. E. and Henze, W., Astrophys. J., 1982, 253, 939–948. 31. Rendtel, J., Staude, J., Innes, D. E., Wilhelm, K. and Gurman, J. B., in A Crossroads for European Solar and Heliospheric Physics (ed. Harris, R. A.), ESA SP-417, 1998, pp. 277–280. 32. Fludra, A., Astron. Astrophys., 1999, 344, L75–L78. 33. Brynildsen, N., Leifsen, T., Kjeldseth-Moe, O., Maltby, P. and Wilhelm, K., Astrophys. J., 1999, 511, L121–L124. 34. Bard, S. and Carlsson, M., in Fifth SOHO Workshop (ed. Wilson, A.), ESA SP-404, 1997, pp. 189–191. ACKNOWLEDGEMENTS. The SUMER project is financially supported by DLR, CNES, NASA, and the ESA PRODEX Programme (Swiss contribution). SUMER is a part of SOHO, the Solar and Heliospheric Observatory, of ESA and NASA. Anita Mohan is supported by the DST, New Delhi, under the young scientist programme. CURRENT SCIENCE, VOL. 77, NO. 11, 10 DECEMBER 1999