Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Strongly Coupled Plasma Physics
Strongly Coupled Plasma Physics
Strongly Coupled Plasma Physics
Ebook1,441 pages

Strongly Coupled Plasma Physics

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Charged particles in dense matter exhibit strong correlations due to the exchange and Coulomb interactions, and thus make a strongly coupled plasma. Examples in laboratory and astrophysical settings include solid and liquid metals, semiconductors, charged particles in lower dimensions such as those trapped in interfacial states of condensed matter or beams, dense multi-ionic systems such a superionic conductors and inertial-confinement-fusion plasmas . The aim of the conference was to elucidate the various physical processes involved in these dense materials. The subject areas covered include plasma physics, atomic and molecular physics, condensed matter physics and astrophysics.

LanguageEnglish
Release dateDec 2, 2012
ISBN9780444597595
Strongly Coupled Plasma Physics

Related to Strongly Coupled Plasma Physics

Physics For You

View More

Reviews for Strongly Coupled Plasma Physics

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Strongly Coupled Plasma Physics - Elsevier Science

    OPENING ADDRESS

    Ladies and Gentlemen:

    On behalf of the Organizing Committee for the Twenty-Fourth Yamada Conference on Strongly Coupled Plasma Physics, I wish to begin the Opening Address by recalling how this Conference was conceived, prepared, and organized.

    First of all, it is my privilege to express our heartiest thanks to the Yamada Science Foundation and Professor Yasusada Yamada for their total support of the Conference. As I shall explain shortly, Professor Yamada was essential in the realization of this Conference from the very beginning of its preparation; he supported the idea of having this sort of international conference in Japan and put forward such a proposal to the Board of Directors, of which he himself is a Member.

    Internationally, a proposal for a scientific meeting on strongly coupled plasma physics in Japan this year arose at the closing of the Strongly Coupled Plasma Physics Meeting held three years ago in the summer of 1986 at the University of California, Santa Cruz. After my return to Japan from that meeting, I initiated an effort in such a direction, along with some of my colleagues here, by submitting a proposal for a three-year Research Grant to the Ministry of Education, Science and Culture, for the specific purpose of preparing for an eventual realization of such an international conference in Japan; we were encouraged by successful funding of this proposal. Soon after the approval of the Grant, the International Advisory Board¹ and the local Preparation Committee were formed to consolidate the organizational structure for the Conference. Subsequently, we contacted Professor Yamada and submitted the proposal for the Conference to the Yamada Science Foundation with his support. When the proposal was approved officially by the Board of Directors of the Foundation in May 1988, the current Organizing Committee¹ was formed. We are, therefore, very grateful for the support and encouragement extended to us from various sources; we are pleased that the conference is now a reality.

    Scientifically, the subject area covered by this Conference is fairly wide, ranging over plasma physics, atomic and molecular physics, condensed matter physics, and astrophysics. Some people may remark that the area to be covered is too broad, but in this age of acute specialization, we believe that this sort of conference has its place in promoting cross fertilization between various disciplines in science. It is the hope of all the Members of the Organizing Committee that the week spent together at the foot of Mount Fuji by Lake Yamanaka will enhance scientific activities, mutual friendship, and the spirit of international cooperation.

    Thank you.

    Setsuo Ichimaru


    ¹see page vii

    WELCOME ADDRESS

    Ladies and Gentlemen:

    On behalf of the Yamada Science Foundation, I would like to extend our hearty welcome to all of you who are participating in the Twenty-Fourth Yamada Conference on Strongly Coupled Plasma Physics, especially to those who have come a long way to Japan from abroad.

    The Yamada Science Foundation develops its activities by giving support to creative projects in the field of basic natural science, providing travel funds for scientists, organizing conferences, promoting international collaboration projects and so forth. Among these, the organization of Yamada Conferences, usually held two or three times a year, is one of the most important activities. As the guiding principles to promote these activities, the Board of Directors of the Foundation puts emphasis on the following three symbolic letter ‘I’s. The first ‘I’ represents ‘International’, the second ‘I’ means ‘Interdisciplinary’ and the third and the most important ‘I’ symbolizes ‘Innovative’.

    In this context, the present Conference is particularly suited for the scope of the Foundation: It is international, having many guests from all over the world; it treats an interdisciplinary field traversing astrophysics, plasma physics, superconductivity, physics of glass states etc., and above all, its subject is one of the most creative and updated areas. For instance, if I understand correctly, the topic is directly related to the basic mechanism of high-Tc superconductivity. I believe that this Conference would fulfill the purpose of our Foundation through most active participation of all of you.

    Another important aspect of this Conference is to provide opportunities to develop new friendships among the participants (particularly among young scientists who are not acquainted with each other) not only through hot scientific discussions but also through heartwarming chattering drinking Yamanashi wine; I expect your most active participation in this aspect too.

    Finally, I would like to express our thanks to the Organizing Committee members for their efforts towards the successful performance of the Conference, choosing such a nice, scenic and wine producing district. I hope that all of you will enjoy the discussions and also relax at the foot of Mount Fuji, a symbol of the beauty of Japan.

    Thank you.

    Yasusada Yamada,     Member, Board of Directors, Yamada Science Foundation

    YAMADA SCIENCE FOUNDATION AND THE SCOPE OF YAMADA CONFERENCES

    The Yamada Science Foundation was established in February 1977 in Osaka through the generosity of Mr. Kiro Yamada. Mr. Yamada was President of Rohto Pharmaceutical Company, Limited, a well-known manufacturer of medicines in Japan. He recognized that creative, unconstrained, basic research is indispensable for the future welfare and prosperity of mankind and he has been deeply concerned with its promotion. Therefore funds for this Foundation were donated from his private holdings.

    The principal activity of the Yamada Science Foundation is to offer financial assistance to creative research in the basic natural sciences, particularly in interdisciplinary domains that bridge established fields. Projects which promote international cooperation are also favored. By assisting in the exchange of visiting scientists and encouraging international meetings, this Foundation intends to greatly further the progress of science in the global environment.

    In this context the Yamada Science Foundation sponsors international Yamada Conferences once or twice a year in Japan. Subjects to be selected by the Foundation should be most timely and stimulating. These conferences are expected to be of the highest international standard so as to significantly foster advances in their respective fields.

    EXECUTIVE MEMBERS OF YAMADA SCIENCE FOUNDATION

    Officers:

    Board of Directors

    Leo ESAKI

    Kenichi FUKUI

    Osamu HAYAISHI

    Noburô KAMIYA

    Takeo NAGAMIYA, Director General

    Shuntaro OGAWA, Standing Director

    Syûzô SEKI

    Tomoji SUZUKI

    Yasusada YAMADA

    Auditors:

    Shigekiyo MUKAI

    Jin-ichi TAKAMURA

    Advisors:

    Shiro AKABORI

    Chapter I

    ASTROPHYSICS

    PHASE TRANSITIONS IN DENSE ASTROPHYSICAL PLASMAS

    H.M. VAN HORN,     Department of Physics and Astronomy, C. E. Kenneth Mees Observatory, and Laboratory for Laser Energetics, University of Rochester, Rochester, NY 14627-0011

    The realization that dense plasmas may freeze, forming crystallized cores in white dwarfs and crusts in neutron stars, initiated an exploration of phase transitions in dense stars. In the intervening 20 years, several other types of phase transitions have been studied. In this review, I shall summarize the advances and identify areas that seem to me particularly interesting and important from the point of view of astrophysics. These include the following: (1) The transition to a glassy state has been proposed as an alternative to crystallization in white dwarfs. (2) The freezing and possible phase separation of binary mixtures, which had been suggested as a possible energy source in very cool C/O white dwarfs, seems unlikely to occur, according to recent calculations. (3) Fe/H phase separation, first suggested as a possibility to account for the solar neutrino problem, almost certainly does not occur in the Sun. It may be a significant effect in the evolution of low mass stars and brown dwarfs, however. (4) H/He phase separation almost certainly does occur in giant planets and provides a significant energy source in these bodies. (5) It has been suggested that the metallization of H may occur via a plasma phase transition. Some recent theoretical work supports this idea, but experimantal tests are still necessary.

    1 INTRODUCTION

    The idea that the plasma in the interior of a dense star can exist in more than one phase was first introduced to astrophysics about 30 years ago. Initially, this concept seems to have been regarded mainly as a technical device to permit calculations of the properties of matter at high densities. Thus, Kirzhnits¹, Abrikosov², and Salpeter³ recognized that a crystalline lattice is the lowest energy state of fully ionized matter at T = 0, and they used this fact in computing the equation of state of dense astrophysical plasmas.

    The pioneering calculations of Brush, Sahlin, and Teller⁴ were the first to explore seriously the fluid/solid phase transition in dense matter. They performed Monte Carlo calculations of the Coulomb interaction energy of the so-called one-component plasma (OCP), a distribution of classical, point ions in a uniform, neutralizing background. They found that for sufficiently large values of the Coulomb coupling parameter Γ = (Ze)²/akBT, where a is the radius of the Wigner-Seitz cell containing a single ion, the system undergoes a spontaneous transition from a fluid phase to a solid phase. This discovery stimulated immediate applications to dense stars. Mestel and Ruderman⁵ used the existence of the high-temperature solid phase to compute the thermal properties of matter in cooling white dwarfs, and Van Horn⁶ showed that white dwarf matter freezes while the star may still be hot enough to be observable. The discovery of pulsars and their interpretation as rotating neutron stars provided another application, as it was realized immediately that the surface layers of these stars would freeze into solid crusts⁷.

    Since this early work, our understanding of phase transitions in dense, astrophysical plasmas has advanced considerably. The purpose of this review is to summarize those advances, concentrating primarily on the astrophysical applications.

    2 SOLIDIFICATION OF DENSE ASTROPHYSICAL PLASMAS

    2.1 Crystallization of White Dwarfs

    Following the seminal investigation by Brush et al.⁴, Hansen⁸ and his coworkers⁹ have made the OCP one of the best-studied models in modern statistical physics. It has now become possible to compute tens of millions of configurations for systems containing thousands of particles, so that thermodynamic averages can be calculated with precision, accurate interparticle correlation functions can be obtained, and various transport processes can be evaluated. The most recent and accurate calculations for the OCP of which I am aware are those by Slattery, Doolen, and DeWitt¹⁰,¹¹, who find that the OCP freezes into a bcc lattice at Γ = 178 ± 1. If crystallization of the plasma is prevented entirely, the OCP can undergo a transition to a Coulomb glass at still higher values of Γ¹² (cf. §2.2. below).

    It is important to recognize that a real dense plasma, such as that found in the core of a white dwarf or the crust of a neutron star, differs from an OCP in one very significant way. The OCP by definition has rigid, uniform background without any physical properties. In a real dense plasma, however, the neutralizing background in which the ions move consists of electrons with very definite physical properties. In particular, the Fermi energy EF of the electrons completely dominates the pressure of the matter, while the ions dominate the thermal properties. The latter are strongly affected by the Coulomb energy ECoul, particularly at low temperatures, where ECoul >> kBT or equivalently Γ >> 1. At sufficiently high densities, where EF >> Ecoul, the electron density is very nearly uniform, so that the OCP provides a good approximation for the properties of the ionic component of the plasma. This is the case in the deep interior of a white dwarf. At lower densities, however, the polarization of the electron background by the ions must be taken into account, and this complicates the calculation of the phase diagram of dense plasmas considerably¹³.

    The first calculations of the evolution of white dwarfs which included the full effects of the fluid/solid phase transition were carried out by Lamb¹⁴,¹⁵. For a 1 M⨀, pure¹²C white dwarf model, they found that core crystallization begins at an age ∼ 10⁹ years, when the star has cooled to a luminosity L ≈ 1.6 × 10−3L⨀ and an effective temperature Teff ≈ 13,000 K. This is indeed sufficiently hot and bright to permit direct observational study. Real white dwarfs are observed to have luminosities as faint as ≈ 10−4.5L⨀ and temperatures as low as ∼ 4500 K. However, they are believed to have cores consisting of a mixture of C and O, rather than pure C, and they generally have masses closer to 0.6M⨀ than to 1.0M⨀. Nevertheless, more recent calculations¹⁶ continue to indicate that crystallizing white dwarfs are in principle observable.

    Crystallization releases the latent heat of fusion associated with the formation of the solid lattice. This is a significant fraction of the thermal energy of the plasma and initially slows the cooling of the star appreciably. Eventually, however, the core temperature T of the white dwarf falls below the Debye temperature ΘD of the lattice. The heat capacity subsequently falls as (Td)³, and the star begins to cool increasingly rapidly. Both of these effects are evident in the theoretical white dwarf luminosity function, the number density of stars (per pc−3) per unit interval in log(L/L⨀). The difficulty of determining the observational luminosity function at very low luminosities has prevented a sufficiently accurate determination of this quantity to test these theoretical predictions with present data, but we expect that data from the Hubble Space Telescope will make this possible in the near future. (See also §||| below).

    2.2 Transition to a Glassy State

    Ichimaru et al.¹² have used an improved HNC scheme to study the OCP at large values of Γ. They found that the second peak in pair correlation function g(r) 210, may be in such an amorphous, glassy state, rather than a crystalline state. More recently, Ogata and Ichimaru¹⁷,¹⁸ have investigated the dynamic evolution of the microstructure in supercooled OCPs using Monte Carlo simulations. They studied four cases of rapid quenches from an equilibrium fluid at Γ = 160 to Γ = 200,300, or 400. Except for the quench to Γ = 200, which formed a supercooled fluid rather than glass, all of these cases relaxed to metastable states (Coulomb glasses) with internal energies lying distinctly below that of the fluid, but above the bcc lattice energy.

    These results are consistent with experience in terrestrial laboratories, where the preparation of a metastable state requires either (i) splat cooling, which provides such a rapid quench that the system has no chance to reach the state of lowest internal energy or else (ii) very gradual cooling through the fluid/solid phase transition without external disturbances.

    Neither of these situations obtains in white dwarfs, however. The white dwarf cooling timescale near the onset of freezing is billions of years, amply long to achieve microscopic equilibrium and avoid the metastable glassy state. Further, white dwarfs are sufficiently noisy that it seems unlikely that a metastable state could remain undisturbed. Surface convection zones, which are present in all cool white dwarfs, provide a noise source intrinsic to the star. In addition, all DA white dwarfs believed to pass through the pulsationally unstable ZZ Ceti phase, where the star undergoes global oscillations. Thus, the conditions necessary for the formation of a metastable, glassy state do not appear to occur in a white dwarf, and I therefore expect them to freeze into a crystalline lattice rather than a glass.

    3 FREEZING OF C/O WHITE DWARFS AND THE AGE OF THE GALAXY

    Until quite recently, the generally accepted view was that our Galaxy had been formed in a rapid inital collapse¹⁹. In this picture, the ages of all the components of the Galaxy must be the same as the ages of the oldest globular clusters, (15 ± 3) × 10⁹ years ≡ 15 ± 3 Gyr²⁰. Recently, however, methods based on nucleocosmochronology and on the cooling ages of the white dwarfs have been introduced, both of which yield ages for the galactic disk that are much younger than this. For example, Malaney and Fowler²¹ have obtained a galactic age tG 12 Gyr from both Th/Nd and Eu/Ba ratios.

    The other new method of obtaining the age of the galactic disk, introduced by Winget et al.²², combines the observed deficiency of white dwarfs having luminosities L < 10−4.5L⨀²³ with white dwarf cooling theory. The result gives tG ≈ 9 Gyr, consistent with the results of nucleocosmochronology but in conflict with the ages of the globular clusters. A more recent and completely independent calculation by Iben and Laughlin²⁴ obtains a disk age ∼ 9 Gyr, in agreement with the results of Winget et al.²².

    Can carbon-oxygen phase separation explain this discrepancy? Until recently, it had been thought that separation of the C/O plasma in a white dwarf into C-rich and O-rich phases upon freezing, as first proposed Stevenson²⁵, might account for the difference between the white dwarf cooling age and the ages of the globular clusters. As shown by Mochkovitch²⁶, the sinking of the denser, O-rich solid releases substantial gravitational energy, slowing the cooling of a white dwarf and lengthening its age. If this were to occur, the disk of the Galaxy could be much older than the current estimate given by the white dwarf chronometer.

    Lengthening the age of a cool white dwarf also increases the luminosity function at these faint magnitudes, however. Garcia-Berro et al.²⁷ have recently computed the effect of complete phase separation upon the luminosity function, and they find a large and potentially observable effect. Thus, if phase separation were to occur, it could eventually be subjected to observational test.

    It now seems very unlikely that C/O phase separation will take place, however. Barrat et al.²⁸ have recomputed the phase diagram for a C/O mixture and have found it to be of the spindle type rather than the eutectic type suggested by Stevenson. They conclude that significant phase separation does not occur, and that the maximum increase in the white dwarf ages associated with the freezing of this binary plasma is about 0.5 Gyr. Still more recently, Ichimaru, lyetomi, and Ogata²⁹ have also recomputed the C/O phase diagram and found it to be of an azeotropic form. Like Barrat et al., they have concluded that significant phase separation does not occur and that the effect on white dwarf ages is minimal. These two studies essentially close the book on this effect.

    The most attractive possibility for resolving the difference between the ages obtained from white dwarf cooling and nucleocosmochronology, on the one hand, and from cluster ages, on the other, seems to me to be to abandon the hypothesis that the disk of the Galaxy is the same age as the halo. Indeed, according to Norris and Green³⁰, There seems no compelling reason to believe that the Galactic disk in the solar neighborhood has any major stellar component as old as the disk globular clusters. They favor a more gradual formation process and point out that the pressure-supported collapse models of Larson³¹ are in best accord with observations. In Larson’s model 6, after ∼ 2 Gyr, the disk is confined to within ∼ 5 kpc of center. Only after a further several Gyr does the disk form at the solar distance from the center. They regard this as the most natural explanation of the apparent relative youth of the disk in the solar neighborhood. Larson³², too, advocates this solution.

    In the end, the definitive determination of the age of the galactic disk will almost certainly require the Hubble Space Telescope. In preparation for the launch of this instrument, Tamanaha et al.³³ have recently undertaken a calculation of the contribution to the white dwarf luminosity function of the stars which completed their evolution during the formation of the galactic halo. Such a possibility was first discussed by Larson³⁴ who postulated bimodal star formation, with a high-mass star formation mode occurring preferentially in the early history of the galactic disk.

    Tamanaha et al.³³ point out that after ∼ 15 Gyr, a 0.8 M⨀ white dwarf will have cooled to L ∼ 10−6L⨀, implying MV ∼ +20, just below the current limit of detection. Though there exist increasing uncertainties in the microphysics at lower luminosities, these authors have made a first effort to explore this domain. Their most interesting finding is that the most extreme models do allow the entire halo dark matter to consist of white dwarfs, with ages thalo = 12 to 13 Gyr. Younger halos cannot supply all the dark matter. This is a testable result, and when the HST is launched, it will surely be one of the priorities for observational investigation.

    4 THE POSSIBILITY OF Fe/H PHASE SEPARATION IN LOW MASS STARS

    The cumulative result from the ³⁷CI experiment which has been running in the Homestake Mine for the past two decades yields a solar neutrino production rate of 2.0 ± 0.3 SNU³⁵. For comparison, the most recent value from the standard solr model is 7.9(1 ± 0.33) SNU³⁶. This discrepancy has stimulated numerous theoretical efforts to find a solution³⁶,³⁷.

    One suggestion for the resolution of this problem was the hypothesis that the center of the Sun may be enriched in heavy elements relative to the abundances detected at the solar surface. A solar model with a small, inert Fe core, formed by draining Fe out of the surrounding H-burning region, would have a significantly different thermal structure from the standard model, perhaps yielding a substantially lower neutrino emission rate.

    Motivated by this idea, Pollock and Alder³⁸ were led to consider the possibility that Fe may undergo phase separation from the H-rich plasma at the high pressures in the solar interior. They performed a hypernetted chain (HNC) calculation of the thermodynamic properties of an Fe/H plasma under conditions like those in Sun (T = 1.5 × 10⁷ K, P = 10⁵ Mbar, Fe concentration = 2.5 × 10−5 ionic mole fraction), using a Debye-Hückel model for the interactions. From the resulting Gibbs free energy G, they computed various simplified Fe/H plasma phase diagrams, which they used to study phase separation in this system. Their results suggested that Fe might indeed undergo phase separation under these conditions, but they emphasized that more accurate calculations would be needed to confirm or refute this suggestion.

    This prospect led to several subsequent efforts to carry out more accurate theoretical calculations of the phase diagram of Fe/H plasmas. The most recent of which I am aware is that of lyetomi and Ichimaru³⁹, who have used more accurate calculations for electron-screened ion plasmas in the HNC approximation. They find that screening is a substantial effect, but that it is not enough to produce Fe/H phase separation. They find the critical point for demixing to be given by Tcrit ≈ 5.5 × 10⁶ K, xcrit ≈ 1.7 × 10−2 at P = 10⁵ Mbar, where x = NFe/(NH + NFe). As the temperature at the critical point is only about 1/3 the value at the center of the Sun, lyetomi and Ichimaru conclude that phase separation does not occur in Sun and thus cannot resolve the solar neutrino problem.

    The conclusion that the Sun is too hot for Fe/H phase separation does not exclude other interesting astrophysical possibilities, however. For example, this may occur in very low-mass main sequence stars, which have significantly lower internal temperatures than does the Sun. This possibility has not yet been explored, but preliminary estimates lend credibility to this idea. In their detailed study of the evolution of very low mass stars, D’Antona and Mazzitelli⁴⁰,⁴¹ found Tc ≈ 4 × 10⁶ K and ρc ≈ 600 g cm−3 at the center of a 0.1 M⨀ H-burning main sequence star. This temperature is well below the critical temperature calculated by lyetomi and Ichimaru³⁹, and as the density is about four times greater than that at the center of the Sun, it seems quite likely that Fe/H phase separation will occur in such low-mass stars. Further calculations, under conditions appropriate to such objects, are clearly essential.

    Brown dwarf stars, which become degenerate and cease gravitational contraction before they become hot enough to ignite H-burning, are still lower in temperature than low-mass main sequence stars. They are thus even more promising candidates for Fe/H phase separation⁴². For example, in a 0.05 M⨀ = 50 MJ brown dwarf, where MJ is the mass of the planet Jupiter, D’Antona and Mazzitelli⁴⁰ find a central temperature Tc ≈ 2 × 10⁶ K and a central density ρc ≈ 450 g cm−3 at an age of about 2 × 10⁹ years. In this object, the temperature has already begun to drop, and – as the temperature already appears to be below the critical temperature for Fe/H phase separation – it seems inescapable that phase separation must be taking place.

    5 H/He PHASE SEPARATION IN GIANT PLANETS

    The discovery that Jupiter radiates significantly more energy than it receives from the Sun triggered several theoretical efforts to find an explanation. Smoluchowski⁴¹ seems to have been the first to discuss H/He phase equilibrium in Jupiter and to relate phase separation to the excess luminosity observed. Salpeter⁴⁴ subsequently pointed out that H/He phase separation can occur even in the presence of the vigorous convection present throughout the interior of Jupiter, with the denser He raining out to form a He-enriched core. The associated release of gravitational energy provides a substantial energy source, which greatly lengthens the cooling time. Salpeter also pointed out the possibility of H/He phase separation as an energy source in low-mass main sequence stars and brown dwarfs, but to my knowledge, no detailed evolutionary calculations have yet been carried out incorporating this energy source.

    Stevenson and Salpeter⁴⁵ have carried out a comprehensive study of the H/He phase diagram, using Stevenson’s⁴⁶ detailed computation of the Gibbs free energy for this system. For a mixture with a solar H/He ratio (10% He by number), these calculations predict a miscibility gap in fluid He/metallic H mixtures at Mbar pressures for T 10⁴ K. MacFarlane⁴⁷, however, finds from a Thomas-Fermi-Dirac calculation that the critical temperatures are much lower than those obtained by Stevenson and Salpeter, so the issue is currently unresolved.

    Stevenson and Salpeter1 Gyr ago). This in turn implies that the critical temperature for the postulated molecular-to-metallic H phase transition (the plasma phase transition ≡ PPT; cf. §VI below) cannot exceed ∼ 20,000 K. If there is no first-order PPT, the phase diagrams predict H/He phase separation, with the denser He droplets raining out, when the internal temperature of the planet falls sufficiently low. For a reasonable choice of parameters, this occurs within the domain of the giant planets. Curiously, the core temperatures may actually increase during phase separation, even though the surface temperature decreases monotonically with increasing age. Alternatively, if the PPT does occur, phase separation may proceed by the formation of He-poor bubbles of metallic H, which rise buoyantly and enrich the overlying layers. In either case, H/He phase separation releases sufficient gravitational energy to increase the cooling time of a giant planet by as much as a factor of four or five.

    Saturn, like Jupiter, radiates more energy than it receives from the Sun. The Voyager 1 flyby showed this factor to be 1.78 ± 0.009⁵⁰. Unlike the case of Jupiter, however, homogeneous evolutionary models for Saturn reach its current luminosity after only ∼ 2 Gyr, rather than the 4.5 Gyr age of the solar system. Thus, H/He phase separation with the accompanying release of gravitational energy appears to be essential to explain the observed excess luminosity⁴⁸.

    Hubbard and Stevenson⁵¹ have estimated the gravitational energy release from H/He phase separation in Saturn to be ∼ (1.5 ± 0.7) × 10¹² erg (g – He)−1. From this they estimate the time since the onset of He differentiation to be ∼ 2 × 10⁹ years. Again, detailed evolutionary calculations, including the effects of phase separation, have not yet been done.

    6 THE METALLIZATION OF H

    Since Wigner and Huntington⁵² first pointed out that H must become a monatomic metal at high densities, there has been considerable interest in the metallization of this element. A thorough review of the literature on these topics is beyond the scope of this paper, but significant advances have occurred recently, which I must at least mention. A good summary of the theoretical and experimental situation at high pressures has recently been provided by Mao and Hemley⁵³.

    6.1 The Metallization of Hydrogen at Low Temperatures

    Immediatiely following Wigner and Huntington’s⁵² prediction of the existence of metallic H, Wildt⁵⁴ and Critchfield⁵⁵ pointed out that this would have consequences for the internal structures of the giant planets. A rather thorough early discussion of the phase diagram of H and the transition from molecular H2 to metallic H was given by de Marcus⁵⁶, who located the transition pressure between 1.93 and 3.5 Mbar, and who applied the results to construct detailed models of Jupiter and Saturn.

    In the 1970s, various experimental investigations of the properties of compressed H became possible, at pressures approaching those anticipated for the transition to the metallic phase. Shock tube experiments had extended up to pressures of 760 kbar⁵⁷,⁵⁸. Some claims had been made that metallization of H had been actually achieved experimentally⁴⁵,⁴⁸,⁵³, but these results are generally discounted because of questions about the reliability of the experiments. A second major group of high pressure experiments has utilized the diamond anvil cell⁵⁹. Ross⁶⁰ has recently summarized both static, diamond anvil-cell measurements and shock measurements on dense hydrogen, and he estimates the presumed first-order molecular-to-metallic phase transition to occur between 3.1 to 3.6 Mbar. Since then, the diamond anvil cell experiments have achieved pressures in excess of 200 GPa (≡ 2 Mbar) and have produced quite exciting new results, which I summarize below⁶¹.

    The issue for the physics of compressed H is the exact nature of the mechanism by which the transition to the metallic phase occurs. Does it happen as a first-order phase transition or through band-overlap⁶²?

    Friedli and Ashcroft⁶² carried out one of the first detailed band-structure calculations for compressed molecular H2, assuming it to be in the Pa3 phase. Their calculations were done for a static lattice, but they recognized the importance of the zero-point motion of the very light protons. They found the band gap to vanish at rs = 1.48, where the highest valence band crosses the lowest conduction band, and corresponds to a second-order metal-insulator transition. Chakravarty et al.⁶² subsequently performed a density-functional calculation of the equation of state for static lattices of both molecular and metallic H. They conclude that remnant molecular pairing is preferred in the band-overlap metallic state; i.e. the metallic state is molecular. In their calculations, the metal-insulator transition occurs at ∼ 2 Mbar, while complete dissociation occurs only at very high densities (rs ∼ 1.1). A completely different approach was taken by Cepperley and Alder64, who did quantum Monte Carlo calculations of the energy of light elements at T = 0. Application of their method to hydrogen yields a molecular H2 to monatomic metallic H transition at 2.8 Mbar.

    More recently, Min, Jansen, and Freeman⁶⁵ have computed the band structure for solid molecular H2 in the Pa3 lattice structure, with results very similar to those obtained by Freidli and Ashcroft⁶³. Min et al. find two pressure-induced transitions. (1) At 1.7 ± 0.2 Mbar they obtain an insulator-to-metal transition within solid molecular H2, which occurs by a combination of band-overlap and bond-length relaxation. This is a second-order electronic transition that occurs without a corresponding structural rearrangement. (2) At 4± 1 Mbar Min et al. find a first-order structural phase transition to the monatomic metallic hcp phase.

    Experimental support for some of these ideas has been obtained quite recently in very high pressure diamond anvil cell experiments. Hemley and Mao⁶⁵ have discovered that solid H2 undergoes a structural phase transition at 145 GPa, as evidenced by an abrupt discontinuity in the intramolecular vibron frequency. They initially interpreted this as the orientational ordering transition which is expected to occur at some pressure. Ashcroft⁶⁷ has pointed out that the issue of whether or not rotation of the H2 molecule is hindered is critical to understanding the nature of the high pressure transitions. If the H2 rotation is stopped, it cannot prevent band overlap, and an abrupt transition to the metallic state occurs. The absence of such an abrupt onset of metallic properties thus implies that hindered rotation is not the explanation for the Carnegie experiments. Conversely, rotation of the H2 molecule does prevent band overlap.

    More recently, Barbee et al.⁶⁸ have computed the total energy for several different lattice structures, using a plane-wave basis extending up to 36 Ry. They included the important proton zero-point motion in the quasiharmonic approximation. For P < 50 Gpa, they found the enthalpies of the Pa3 and the m − hcp structures to be indistinguishable within their limits of error. For P > 80 ± 15 GPa, however, the m − hcp structure is more stable, in agreement with experiments⁶⁹,⁷⁰. The H2 molecules are already oriented (rotationally hindered) in this pressure range, according to their calculations. There is no phase transition associated with this change, because the molecules can orient parallel with the hexagonal axis of this structure without changing the symmetry. Thus, they believe that the phase transition observed by Hemley and Mao⁶⁶ at ∼ 150 GPa is not associated with molecular orientation. For P > 380 ± 50 Gpa, Barbee et al. find a highly anisotropic, filamentary, primitive-hexagonal structure to be the most stable. The transition to the monatomic bcc phase occurs at 860 ± 100 Gpa. They also note that a metal-insulator transition is expected near 200 GPa in the m − hcp phase, but their method of calculation is unable to predict this accurately.

    The most recent experiments by Mao and Hemley⁵³ may finally have detected hydrogen metallization. In a set of seven experiments, they found the intensity of Raman scattering of the molecular vibron of H2 to appear first near 180 GPa, suggesting the onset of electronic transitions near the laser excitation energy (2.54 eV). At higher pressure, the resonance frequency shifts to longer wavelengths, implying a redshift of the electronic transitions. The authors emphasize that the change of properties with pressure is not discontinuous. Above ∼ 150 GPa, light transmission through the part of the hydrogen sample at the highest pressure within the cell gradually decreased, suggesting that the electronic excitation threshold had decreased from the zero-pressure value of 10.9 eV into the visible region. Reflectivity measurements also show an increase above 150 GPa. The lack of a sharp absorption edge may indicate the closure of an indirect gap; this depends upon the exact crystal structure, which is not yet known experimentally for this pahse. Mao and Hemley believe that the maximum P achieved in this study was > 250 GPa. There is some evidence that dissociation may already have occurred occurred at the highest pressures, but the results are not yet conclusive.

    Thus, theory and experiment seem to be converging, and our understanding of the pressure-induced transitions of hydrogen has improved dramatically in just the past few years.

    6.2 The Plasma Phase Transition in Dense H

    Considerations of the transition to the fully ionized, dense H plasma at elevated temperature have been motivated in part by an interest in computing the opacity of matter at high temperatures and densities⁷¹ and in part by interest in the nature of the phase diagram⁷². The opacity-motivated research, biased by the knowledge that thermal ionization is a continuous process, has focused on rather ad hoc modifications of the Saha equation which lead to pressure-ionization through lowering of the ionization potential⁷³. Conversely, research on the phase diagram of H at high pressures and temperatures has been stimulated by the availability of shock-wave experiments capable of reaching the regime where pressure-ionization becomes important⁵⁷.

    In their detailed study of the H/He phase diagram, Stevenson and Salpeter10⁴ K. If this is true, the coexistence curve must end in a second critical point (the first critical point marking the termination of the gas-liquid coexistence curve), because ionization is a continuous process at lower pressures. The model calculation of Stevenson and Salpeter assumes a first-order molecular H2-to-monatomic metallic H phase transition and predicts the second critical point to lie at Tc = 3500 K, Pc ≈ 3 Mbar.

    The next step was taken by Robnik and Kundt⁷⁴, who computed a model for the free energy of a mixture of protons, electrons, and H atoms, but with no H2 molecules. They ignored electron energy band effects (which Stevenson⁴⁶ had included) and the interactions between the H atoms and the charged particles. With these approximations, they found the second critical point to lie at Tc = 19,000 K, Pc = 24 GPa = 0.24 Mbar. They also pointed out the connection between this plasma phase transition (PPT) and the molecular-metallic hydrogen phase transition. Ebeling and Richert⁷⁵ (1985) subsequently performed an independent calculation of the coexistence curve and second critical point. They also found a dramatic lowering of the pressure along the neutral/ionized coexistence curve for T > 2000 K, with the location of the second critical point given by Tc = 16,500 K, Pc = 22.8 GPa, ρc = 0.13g cm−3. Marley and Hubbard⁷⁶,⁷⁷ next computed the PPT by equating the chemical potential computed for the molecular phase to that of the metallic phase. For molecular H2, they used a model free energy fitted to the shock-wave experiments of Nellis et al.⁵⁷, while for the metallic phase they used the free energy given by Hubbard and DeWitt¹³. Their method of calculation does not yield a critical point. Interestingly, like the calculations by Robnik and Kundt⁷⁴ and by Ebeling and Richert⁷⁵, Marley and Hubbard find dT/dP > 0 along the coexistence curve, implying ΔS < 0 at the presumed first-order phase transition. Thus, the latent heat associated with this transition is negative.

    The most recent contribution to this discussion is a new calculation by Saumon and Chabrier⁷⁸. They have computed a new equation of state for fluid hydrogen, based on sophisticated models for the free energies of the neutral and ionized states. For the low-density, low-temperature, neutral state, the model free energy is computed from hard-sphere fluid perturbation theory for a mixture of H2 molecules and H atoms. The model treats the internal states of the atoms and molecules using the occupation probability formalism recently developed for opacity calculations by Hummer and Mihalas⁷⁹. For the high-temperature or high-density, fully ionized phase (kT 2 g cm−3, corresponding to rs 1), the plasma is treated as a linearly screened ionic fluid plus a partially degenerate electron liquid, using an accurate fit to the exchange and correlation energy⁸⁰.

    In the region of partial ionization, Saumon and Chabrier’s two separate calculations are combined into a single, unified model for the free energy. In this regime, interactions between neutral and charged species are included through a polarization potential. The resulting free energy is minimized to obtain the relative abundances of H2, H, H+, and e−, with the requirement of electrical neutrality imposed. This calculation predicts, rather than assumes a first-order plasma phase transition. The second critical point is found to lie at Tc = 15,000 K, Pc = 0.646 Mbar = 64.6 GPa, and ρc = 0.36g cm−3. The latent heat is again found to be negative, there is a substantial density discontinuity across the phase boundary, and molecular dissociation occurs simultaneously with pressure- ionization. In contrast to some previous work, molecules are found to be the dominant species in the neutral phase at high density.

    While it is the most detailed calculation of the plasma phase transition to date, Saumon and Chabrier’s result still employs some potentially important approximations. For example, they neglect band structure effects in the dense fluid. Thus, we still need experiments to tell us the real facts. Fortunately, the new calculations keep the possibility of experimental test within reach in the near future.

    7 SUMMARY AND CONCLUSIONS

    In the past few decades, phase transitions have assumed an increasingly important role in astrophysics. We now believe that the centers of the coolest white dwarfs and the surface layers of neutron stars freeze solid as they cool, although they remain fully ionized. The apparent lack of C/O phase separation during crystallization of the cores of white dwarfs has led us to assign an age ∼ 9 Gyr to the disk of our Galaxy and is helping to bring about a new view of the process by which the Galaxy formed. Phase separation of Fe from H, originally conceived as a possible solution to the solar neutrino problem, seems not to occur in the Sun but almost certainly occurs in low-mass stars and brown dwarfs. H/He phase separation seems essential to explain the excess IR luminosity of Saturn. Quite recently, striking new evidence has been obtained in support of the band-overlap model for the metallization of hydrogen, which may have consequences for models of the interiors of the giant planets, and new theoretical calculations support the existence of the plasma phase transition in hydrogen at high densities.

    Many interesting problems remain to be attacked. For example, I have not even discussed several other types of phase transitions, including the possible formation of a pion condensate in the dense interior of a neutron star⁸⁰, the possibility of formation of a neutron crystal in the centers of massive neutron stars⁸¹, or the suggested existence of phase transitions in dense neutron star matter, from nuclei to bubbles and from bubbles to uniform matter⁸².

    In addition, we need to carry out detailed evolutionary calculations for low-mass stars and giant planets, which include the full effects of Fe/H and H/He phase separation. The H/He phase diagram must first be recomputed with sufficient accuracy to give accurate values for the critical point for H/He phase separation, however. We also need to explore the possibility of Fe/C phase separation as an energy source in cooling white dwarfs⁴⁴.

    Metallic hydrogen remains a stimulating challenge to experimentalists and theorists alike. Has the monatomic metllic state finally been formed⁵³? Does metallic H solidify at all at low temperatures⁸³? Is metallic H a superconductor, either in the solid phase, if it exists⁸⁴,⁸⁵, or in the liquid phase⁸⁶,⁸⁷? Does the plasma phase transition exist at high temperatures?

    Clearly there will be enough challenges to keep us all happily occupied for years to come.

    ACKNOWLEDGEMENTS

    I am grateful to Neil Ashcroft, Gilles Chabrier, Bill Forrest, David Stevenson, Hugh DeWitt, Marvin Ross, Didier Saumon, and David Young for numerous helpful discussions. This work has been supported in part by the National Science Foundation under grants PHY 88-08146 and AST 88-20322 and in part by the National Aeronautics and Space Administration under grant NAGW- 1476, all through the University of Rochester.

    REFERENCES

    1. Kirzhnits, D.A. Soviet Phys. - J. E. T. P. 1960; 11:365.

    2. Abrikosov, A.A. Soviet Phys. - J. E. T. P. 1961; 12:1254.

    3. Salpeter, E.E. Ap. J. 1961; 134:669.

    4. Brush, S.G., Sahlin, H.L., Teller, E. J. Chem. Phys. 1966; 45:2102.

    5. Mestel, L., Ruderman, M.A. M. N. R. A. S. 1967; 136:27.

    6. Van Horn, H.M. Ap. J. 1968; 151:227.

    7. Ruderman, M.A. Nature. 1968; 218:1129.

    8. Hansen, J.P. Phys. Rev. A. 1973; 8:3096.

    9. Hansen, J.P. J. de Phys. 1979; 41:C2–C3. [Coll. C2].

    10. Slattery, W.L., Doolen, G.D., DeWitt, H.E. Phys. Rev A. 1980; 21:2087.

    11. Slattery, W.L., Doolen, G.D., DeWitt, H.E. Phys. Rev. A. 1982; 26:2255.

    12. Ichimaru, S., Iyetomi, H., Mitake, S., Itoh, N. Ap. J. 1983; 265:L83.

    13. Hubbard, W.B., DeWitt, H.E. Ap. J. 1985; 290:388.

    14. Lamb, D. Q., Ph.D. thesis, University of Rochester. 1974.

    15. Lamb, D.Q., Van Horn, H.M. Ap. J. 1975; 200:306.

    16. Wood, M. A., Ph.D. thesis, University of Texas (Austin) 1989.

    17. Ogata, S., Ichimaru, S. Phys. Rev. A. 1989; 39:1333.

    18. Ogata, S., Ichimaru, S. J. Phys. Soc. Jap. 1989; 58:356.

    19. Eggen, O.J., Lynden-Bell, D., Sandage, A. Ap. J. 1962; 136:748.

    20. Iben, I., Renzini, A. Phys. Rep. 1984; 105:330.

    21. Malaney, R.A., Fowler, W.A. M. N. R. A. S. 1989; 237:67.

    22. Winget, D.E., Hansen, C.J., Liebert, J.W., Van Horn, H.M., Fontaine, G., Nather, R.E., Kepler, S.O., Lamb, D.Q. Ap. J. 1987; 315:L77.

    23. Liebert, J., Dahn, C.C., Monet, D.G. Ap. J. 1988;

    Enjoying the preview?
    Page 1 of 1