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

Only $11.99/month after trial. Cancel anytime.

Readings from the Treatise on Geochemistry
Readings from the Treatise on Geochemistry
Readings from the Treatise on Geochemistry
Ebook2,160 pages16 hours

Readings from the Treatise on Geochemistry

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Readings from the Treatise on Geochemistry offers an interdisciplinary reference for scientists, researchers and upper undergraduate and graduate level geochemistry students that is more affordable than the full Treatise. For professionals, this volume will provide an overview of the field as a whole. For students, it will provide more in-depth introductory content than is found in broad-based geochemistry textbooks. Articles were selected from chapters across all volumes of the full Treatise, and include: The Origin and Earliest History of the Earth, Compositional Evolution of the Mantle, Evolution of Sedimentary Rocks, Soil Formation, Geochemistry of Groundwater, Geologic History of Seawater, Hydrothermal Processes, and Biogeochemistry of Primary Production in the Sea.

  • Comprehensive, interdisciplinary and authoritative content selected by leading subject experts
  • Robust illustrations, figures and tables
  • Affordably priced sampling of content from the full Treatise on Geochemistry
LanguageEnglish
Release dateMay 25, 2010
ISBN9780123813923
Readings from the Treatise on Geochemistry

Related to Readings from the Treatise on Geochemistry

Related ebooks

Earth Sciences For You

View More

Related articles

Reviews for Readings from the Treatise on Geochemistry

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

    Readings from the Treatise on Geochemistry - Heinrich D Holland

    Readings from the Treatise on Geochemistry

    First Edition

    H.D. Holland

    Harvard University, Cambridge, MA, USA

    University of Pennsylvania, Philadelphia, PA

    K.K. Turekian

    Yale University, New Haven, CT, USA

    Academic Press

    AMSTERDAM • BOSTON • HEIDELBERG • LONDON • NEW YORK • OXFORD

    PARIS • SAN DIEGO • SAN FRANCISCO • SINGAPORE • SYDNEY • TOKYO

    Table of Contents

    Cover image

    Title page

    Copyright page

    Introduction

    Contributors

    1: Origin of the Elements

    1.1 INTRODUCTION

    1.2 ABUNDANCES AND NUCLEOSYNTHESIS

    1.3 INTERMEDIATE MASS STARS: EVOLUTION AND NUCLEOSYNTHESIS

    1.4 MASSIVE-STAR EVOLUTION AND NUCLEOSYNTHESIS

    1.5 TYPE Ia SUPERNOVAE: PROGENITORS AND NUCLEOSYNTHESIS

    1.6 NUCLEOSYNTHESIS AND GALACTIC CHEMICAL EVOLUTION

    2: The Origin and Earliest History of the Earth

    2.1 INTRODUCTION

    2.2 OBSERVATIONAL EVIDENCE AND THEORETICAL CONSTRAINTS PERTAINING TO THE NEBULAR ENVIRONMENT FROM WHICH EARTH ORIGINATED

    2.3 THE DYNAMICS OF ACCRETION OF THE EARTH

    2.4 CONSTRAINTS FROM LEAD AND TUNGSTEN ISOTOPES ON THE OVERALLTIMING, RATES, AND MECHANISMS OF TERRESTRIAL ACCRETION

    2.5 CHEMICAL AND ISOTOPIC CONSTRAINTS ON THE NATURE OF THE COMPONENTS THAT ACCRETED TO FORM THE EARTH

    2.6 EARTH’S EARLIESTATMOSPHERES AND HYDROSPHERES

    2.7 MAGMA OCEANS AND CORE FORMATION

    2.8 THE FORMATION OF THE MOON

    2.9 MASS LOSS AND COMPOSITIONAL CHANGES DURING ACCRETION

    2.10 EVIDENCE FOR LATE ACCRETION, CORE FORMATION, AND CHANGES IN VOLATILES AFTER THE GIANT IMPACT

    2.11 THE HADEAN

    2.12 CONCLUDING REMARKS—THE PROGNOSIS

    ACKNOWLEDGMENTS

    3: Sampling Mantle Heterogeneity through Oceanic Basalts: Isotopes and Trace Elements

    3.1 INTRODUCTION

    3.2 LOCAL AND REGIONAL EQUILIBRIUM REVISITED

    3.3 CRUST–MANTLE DIFFERENTIATION

    3.4 MID-OCEAN RIDGE BASALTS: SAMPLES OF THE DEPLETED MANTLE

    3.5 OCEAN ISLAND, PLATEAU, AND SEAMOUNT BASALTS

    3.6 THE LEAD PARADOX

    3.7 GEOCHEMICAL MANTLE MODELS

    ACKNOWLEDGMENTS

    4: Compositional Model for the Earth’s Core

    4.1 INTRODUCTION

    4.2 FIRST-ORDER GEOPHYSICS

    4.3 CONSTRAINING THE COMPOSITION OF THE EARTH’S CORE

    4.4 A COMPOSITIONAL MODEL FOR THE CORE

    4.5 RADIOACTIVE ELEMENTS IN THE CORE

    4.6 TIMING OF CORE FORMATION

    4.7 NATURE OF CORE FORMATION

    4.8 THE INNER CORE, ITS CRYSTALLIZATION, AND CORE-MANTLE EXCHANGE

    4.9 SUMMARY

    ACKNOWLEDGMENTS

    5: Composition of the Continental Crust

    5.1 INTRODUCTION

    5.2 THE UPPER CONTINENTAL CRUST

    5.3 THE DEEP CRUST

    5.4 BULK CRUST COMPOSITION

    5.5 IMPLICATIONS OF THE CRUST COMPOSITION

    5.6 EARTH’S CRUST IN A PLANETARY PERSPECTIVE

    5.7 SUMMARY

    ACKNOWLEDGMENTS

    6: The History of Planetary Degassing as Recorded by Noble Gases

    6.1 INTRODUCTION

    6.2 PRESENT-EARTH NOBLE GAS CHARACTERISTICS

    6.3 BULK DEGASSING OF RADIOGENIC ISOTOPES

    6.4 DEGASSING OF THE MANTLE

    6.5 DEGASSING OF THE CRUST

    6.6 MAJOR VOLATILE CYCLES

    6.7 DEGASSING OF OTHER TERRESTRIAL PLANETS

    6.8 CONCLUSIONS

    ACKNOWLEDGMENT

    7: Soil Formation

    7.1 INTRODUCTION

    7.2 FACTORS OF SOIL FORMATION

    7.3 SOIL MORPHOLOGY

    7.4 MASS BALANCE MODELS OF SOIL FORMATION

    7.5 PROCESSES OF MATTER AND ENERGY TRANSFER IN SOILS

    7.6 SOIL DATA COMPILATIONS

    7.7 CONCLUDING COMMENTS

    8: Global Occurrence of Major Elements in Rivers

    8.1 INTRODUCTION

    8.2 SOURCES OF DATA

    8.3 GLOBAL RANGE OF PRISTINE RIVER CHEMISTRY

    8.4 SOURCES, SINKS, AND CONTROLS OF RIVER-DISSOLVED MATERIAL

    8.5 IDEALIZED MODEL OF RIVER CHEMISTRY

    8.6 DISTRIBUTION OF WEATHERING INTENSITIES AT THE GLOBAL SCALE

    8.7 GLOBAL BUDGET OF RIVERINE FLUXES

    8.8 HUMAN ALTERATION OF RIVER CHEMISTRY

    8.9 CONCLUSIONS

    9: Geochemistry of Groundwater

    9.1 INTRODUCTION

    9.2 GROUNDWATER GEOCHEMISTRY IN SOME COMMON GEOLOGIC SETTINGS

    9.3 REDUCTION/OXIDATION PROCESSES

    9.4 CONTAMINANT GEOCHEMISTRY

    9.5 SUMMARY AND CONCLUSIONS

    10: The Oceanic CaCO3 Cycle

    10.1 INTRODUCTION

    10.2 DEPTH OF TRANSITION ZONE

    10.3 DISTRIBUTION OF CO32− ION IN TODAY’S DEEP OCEAN

    10.4 DEPTH OF SATURATION HORIZON

    10.5 DISSOLUTION MECHANISMS

    10.6 DISSOLUTION IN THE PAST

    10.7 SEDIMENT-BASED PROXIES

    10.8 SHELL WEIGHTS

    10.9 THE BORON ISOTOPE PALEO pH METHOD

    10.10 Zn/Cd RATIOS

    10.11 DISSOLUTION AND PRESERVATION EVENTS

    10.12 GLACIAL TO INTERGLACIAL CARBONATE ION CHANGE

    10.13 NEUTRALIZATION OF FOSSIL FUEL CO2

    11: Hydrothermal Processes

    11.1 INTRODUCTION

    11.2 VENT-FLUID GEOCHEMISTRY

    11.3 THE NET IMPACT OF HYDROTHERMAL ACTIVITY

    11.4 NEAR-VENT DEPOSITS

    11.5 HYDROTHERMAL PLUME PROCESSES

    11.6 HYDROTHERMAL SEDIMENTS

    11.7 CONCLUSION

    12: The Geologic History of Seawater

    12.1 INTRODUCTION

    12.2 THE HADEAN (4.5–4.0 Ga)

    12.3 THE ARCHEAN (4.0–2.5 Ga)

    12.4 THE PROTEROZOIC

    12.5 THE PHANEROZOIC

    12.6 A BRIEF SUMMARY

    ACKNOWLEDGMENTS

    13: Geochemistry of Fine-grained Sediments and Sedimentary Rocks

    13.1 INTRODUCTION

    13.2 CONCEPTUAL MODEL—PROCESSES

    13.3 CONCEPTUAL MODEL: PROXIES

    13.4 GEOCHEMICAL CASE STUDIES OF FINE-GRAINED SEDIMENTS AND SEDIMENTARY ROCKS

    13.5 DISCUSSION: A UNIFIED VIEW OF THE GEOCHEMISTRY OF FINE-GRAINED SEDIMENTS AND SEDIMENTARY ROCKS

    ACKNOWLEDGMENTS

    14: Evolution of Sedimentary Rocks

    14.1 INTRODUCTION

    14.2 THE EARTH SYSTEM

    14.3 GENERATION AND RECYCLING OF THE OCEANIC AND CONTINENTAL CRUST

    14.4 GLOBAL TECTONIC REALMS AND THEIR RECYCLING RATES

    14.5 PRESENT-DAY SEDIMENTARY SHELL

    14.6 TECTONIC SETTINGS AND THEIR SEDIMENTARY PACKAGES

    14.7 PETROLOGY, MINERALOGY, AND MAJOR ELEMENT COMPOSITION OF CLASTIC SEDIMENTS

    14.8 TRACE ELEMENT AND ISOTOPIC COMPOSITION OF CLASTIC SEDIMENTS

    14.9 SECULAR EVOLUTION OF CLASTIC SEDIMENTS

    14.10 SEDIMENTARY RECYCLING

    14.11 OCEAN/ATMOSPHERE SYSTEM

    14.12 MAJOR TRENDS IN THE EVOLUTION OF SEDIMENTS DURING GEOLOGIC HISTORY

    ACKNOWLEDGMENT

    15: Biogeochemistry of Primary Production in the Sea

    15.1 INTRODUCTION

    15.2 CHEMOAUTOTROPHY

    15.3 PHOTOAUTOTROPHY

    15.4 PRIMARY PRODUCTIVITY BY PHOTOAUTOTROPHS

    15.5 EXPORT, NEW AND TRUE NEW PRODUCTION

    15.6 NUTRIENT FLUXES

    15.7 NITRIFICATION

    15.8 LIMITING MACRONUTRIENTS

    15.9 THE EVOLUTION OF THE NITROGEN CYCLE

    15.10 FUNCTIONAL GROUPS

    15.11 HIGH-NUTRIENT, LOW-CHLOROPHYLL REGIONS—IRON LIMITATION

    15.12 GLACIAL–INTERGLACIAL CHANGES IN THE BIOLOGICAL CO2 PUMP

    15.13 IRON STIMULATION OF NUTRIENT UTILIZATION

    15.14 LINKING IRON TO N2 FIXATION

    15.15 OTHER TRACE-ELEMENT CONTROLS ON NPP

    15.16 CONCLUDING REMARKS

    ACKNOWLEDGMENTS

    16: The Contemporary Carbon Cycle

    16.1 INTRODUCTION

    16.2 MAJOR RESERVOIRS AND NATURAL FLUXES OF CARBON

    16.3 CHANGES IN THE STOCKS AND FLUXES OF CARBON AS A RESULT OF HUMAN ACTIVITIES

    16.4 MECHANISMS THOUGHT TO BE RESPONSIBLE FOR CURRENT SINKS OF CARBON

    16.5 THE FUTURE: DELIBERATE SEQUESTERING OF CARBON (OR REDUCTION OF SOURCES)

    16.6 CONCLUSION

    17: Environmental Geochemistry of Radioactive Contamination

    NOMENCLATURE

    17.1 INTRODUCTION

    17.2 THE NATURE AND HAZARDS OF RADIOACTIVE ENVIRONMENTAL CONTAMINATION

    17.3 EXPERIMENTAL AND THEORETICAL STUDIES OF RADIONUCLIDE GEOCHEMISTRY

    17.4 FIELD STUDIES OF RADIONUCLIDE BEHAVIOR

    17.5 APPLICATIONS: DEALING WITH RADIONUCLIDE CONTAMINATION IN THE ENVIRONMENT

    17.6 SUMMARY—CHALLENGES AND FUTURE RESEARCH NEEDS

    ACKNOWLEDGMENTS

    Index

    Copyright

    Elsevier Ltd.

    Radarweg 29, 1043 NX Amsterdam, the Netherlands

    First edition 2010

    Copyright © 2010 Elsevier Ltd. All rights reserved

    Permissions may be sought directly from Elsevier’s Science & Technology Rights

    Department in Oxford, UK: phone (+ 44) (0) 1865 843830; fax (+ 44) (0) 1865 853333; email: permissions@elsevier.com. Alternatively visit the Science & Technology website at www.elsevierdirect.com/rights for further information

    Notice

    No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made

    British Library Cataloguing in Publication Data

    A catalogue record for this book is available from the British Library

    Library of Congress Control Number: 2009937759

    ISBN: 978–0–12–381391–6

    For information on all Elsevier publications

    visit our website at elsevierdirect.com

    Printed and bound in Spain

    09  10  11  12  10  9  8  7  6  5  4  3  2  1

    Introduction

    H.D. Holland; K.K. Turekian, Executive Editors, September 2009

    These readings from the updated first edition of the Treatise on Geochemistry were chosen to serve as supplements for students in General Geochemistry courses, and to introduce professionals to the field of geochemistry. They are only selections, but they span the entire range of geochemistry that is represented in the Treatise.

    We start with a discussion of the origin of the elements that is basic to any further exploration of the geochemistry of the Solar System and the Earth in particular. This introduction is followed by an essay on the use of isotopic studies of meteorites and mantle-derived rocks to arrive at the history of the Earth’s formation and its separation into the crust, mantle, and core. This essay is followed by chapters on the composition of these three major parts of the Earth, and on the processes associated with plate tectonics that appear to have altered significant portions of the mantle over time. These chapters are followed by an essay on the emplacement of the atmosphere and oceans and on the mechanisms by which volatiles move from the Earth’s interior to its surface.

    Earth surface processes transport and redistribute the elements and their compounds. These processes are tracked in chapters on soil formation and on the chemistry of rivers and groundwater. The oceans modify the geochemistry of many chemical species. Their effect is illustrated by chapters on the marine CaCO3 cycle and on the effects of marine hydrothermal systems on the chemistry and the biology of the oceans. Three chapters are devoted to sedimentary rocks, which carry an impressive record of changes in the composition of the atmosphere and of seawater, and in the Earth’s surface environments during its long history. The evolution of these environments has been influenced strongly by the biosphere. Chapters are therefore devoted to the biogeochemistry of primary organic production in the oceans and to the carbon cycle as a whole.

    The cycles of the elements at the Earth’s surface, together with many other parts of the Earth system are strongly affected by humanity. The final chapter of the Readings deals with the environmental geochemistry of radioactive contaminants. It serves as an example of the human impact and as a warning of potential disasters.

    The chapters in this volume were chosen from the large array in the Treatise on Geochemistry. They are only fragments but we hope that they will convey the wide sweep and scope of the field that is geochemistry.

    Contributors

    R. Amundson     University of California, Berkeley, CA, USA

    W.S. Broecker     Lamont–Doherty Earth Observatory, Columbia University, Palisades, NY

    C.R. Bryan     Sandia National Laboratories, Albuquerque, NM, USA

    F.H. Chapelle     US Geological Survey, Columbia, SC, USA

    P.G. Falkowski     Rutgers University, New Brunswick, NJ, USA

    S. Gao     China University of Geosciences, Wuhan, People’s Republic of China and Northwest University, Xi’an, People’s Republic of China

    C.R. German     Southampton Oceanography Centre, Southampton, UK

    A.N. Halliday     University of Oxford, UK

    A. Heger     University of Chicago, IL, USA

    A.W. Hofmann     Max–Plank–Institut für Chemie, Mainz, Germany and Lamont–Doherty Earth Observatory, Columbia University, Palisades, NY, USA

    H.D. Holland

    Harvard University, Cambridge, MA, USA

    University of Pennsylvania, Philadelphia, PA

    R.A. Houghton     Woods Hole Research Center, MA, USA

    T.W. Lyons     University of California, Riverside, CA, USA

    F.T. Mackenzie     University of Hawaii, Honolulu, HI, USA

    W.F. McDonough     University of Maryland, College Park, USA

    M. Meybeck     University of Paris VI, CNRS, Paris, France

    D. Porcelli     University of Oxford, UK

    R.L. Rudnick     University of Maryland, College Park, MD, USA

    B.B. Sageman     Northwestern University, Evanston, IL, USA

    M.D. Siegel     Sandia National Laboratories, Albuquerque, NM, USA

    J.W. Truran, Jr.     University of Chicago, IL, USA

    K.K. Turekian     Yale University, New Haven, CT, USA

    J. Veizer     Ruhr University, Bochum, Germany and University of Ottawa, ON, Canada

    K.L. Von Damm†     University of New Hampshire, Durham, NH, USA

    1

    Origin of the Elements

    J.W. Truran, Jr.; A. Heger    University of Chicago, IL, USA

    1.1 INTRODUCTION   1

    1.2 ABUNDANCES AND NUCLEOSYNTHESIS   2

    1.3 INTERMEDIATE MASS STARS: EVOLUTION AND NUCLEOSYNTHESIS   3

    1.3.1 Shell Helium Burning and ¹²C Production   4

    1.3.2 s-Process Synthesis in Red Giants   4

    1.4 MASSIVE-STAR EVOLUTION AND NUCLEOSYNTHESIS   5

    1.4.1 Nucleosynthesis in Massive Stars   7

    1.4.1.1 Hydrogen burning   7

    1.4.1.2 Helium burning and the s-process   7

    1.4.1.3 Hydrogen and helium shell burning   7

    1.4.1.4 Carbon burning   8

    1.4.1.5 Neon and oxygen burning   8

    1.4.1.6 Silicon burning   8

    1.4.1.7 Explosive nucleosynthesis   8

    1.4.1.8 The p-process   9

    1.4.1.9 The r-process   9

    1.5 TYPE Ia SUPERNOVAE: PROGENITORS AND NUCLEOSYNTHESIS   9

    1.6 NUCLEOSYNTHESIS AND GALACTIC CHEMICAL EVOLUTION   12

    REFERENCES   14

    1.1

    INTRODUCTION

    Nucleosynthesis is the study of the nuclear processes responsible for the formation of the elements which constitute the baryonic matter of the Universe. The elements of which the Universe is composed indeed have a quite complicated nucleosynthesis history, which extends from the first three minutes of the Big Bang through to the present. Contemporary nucleosynthesis theory associates the production of certain elements/isotopes or groups of elements with a number of specific astrophysical settings, the most significant of which are: (i) the cosmological Big Bang, (ii) stars, and (iii) supernovae.

    Cosmological nucleosynthesis studies predict that the conditions characterizing the Big Bang are consistent with the synthesis only of the lightest elements: ¹H, ²H, ³He, ⁴He, and ⁷Li (Burles et al., 2001; Cyburt et al., 2002). These contributions define the primordial compositions both of galaxies and of the first stars formed therein. Within galaxies, stars and supernovae play the dominant role both in synthesizing the elements from carbon to uranium and in returning heavy-element-enriched matter to the interstellar gas from which new stars are formed. The mass fraction of our solar system (formed ~4.6 Gyr ago) in the form of heavy elements is ~ 1.8%, and stars formed today in our galaxy can be a factor 2 or 3 more enriched (Edvardsson et al., 1993). It is the processes of nucleosynthesis operating in stars and supernovae that we will review in this chapter. We will confine our attention to three broad categories of stellar and supernova site with which specific nucleosynthesis products are understood to be identified: (i) intermediate mass stars, (ii) massive stars and associated type II supernovae, and (iii) type Ia supernovae. The first two of these sites are the straightforward consequence of the evolution of single stars, while type Ia supernovae are understood to result from binary stellar evolution.

    Stellar nucleosynthesis resulting from the evolution of single stars is a strong function of stellar mass (Woosley et al.M/M 10), the temperatures realized in their degenerate cores never reach levels at which carbon ignition can occur. Substantial element production occurs in such stars during the asymptotic giant branch (AGB) phase of evolution, accompanied by significant mass loss, and they evolve to white dwarfs of carbon–oxygen (or, less commonly, oxygen–neon) composition. In contrast, the increased pressures that are experienced in the cores of stars of masses M 10M yield higher core temperatures that enable subsequent phases of carbon, neon, oxygen, and silicon burning to proceed. Collapse of an iron core devoid of further nuclear energy then gives rise to a type II supernova and the formation of a neutron star or black hole remnant (Heger et al., 2003). The ejecta of type IIs contain the ashes of nuclear burning of the entire life of the star, but are also modified by the explosion itself. They are the source of most material (by mass) heavier than helium.

    Observations reveal that binary stellar systems comprise roughly half of all stars in our galaxy. Single star evolution, as noted above, can leave in its wake compact stellar remnants: white dwarfs, neutron stars, and black holes. Indeed, we have evidence for the occurrence of all three types of condensed remnant in binaries. In close binary systems, mass transfer can take place from an evolving companion onto a compact object. This naturally gives rise to a variety of interesting phenomena: classical novae (involving hydrogen thermonuclear runaways in accreted shells on white dwarfs (Gehrz et al., 1998)), X-ray bursts (hydrogen/helium thermonuclear runaways on neutron stars (Strohmayer and Bildsten, 2003)), and X-ray binaries (accretion onto black holes). For some range of conditions, accretion onto carbon–oxygen white dwarfs will permit growth of the CO core to the Chandrasekhar limit MCh = 1.4M , and a thermonuclear runaway in to core leads to a type Ia supernova.

    In this chapter, we will review the characteristics of thermonuclear processing in the three environments we have identified: (i) intermediate-mass stars; (ii) massive stars and type II supernovae; and (iii) type Ia supernovae. This will be followed by a brief discussion of galactic chemical evolution, which illustrates how the contributions from each of these environments are first introduced into the interstellar media of galaxies. Reviews of nucleosynthesis processes include those by Arnett (1995), Trimble (1975), Truran (1984), Wallerstein et al. (1997), and Woosley et al. (2002). An overview of galactic chemical evolution is presented by Tinsley (1980).

    1.2 ABUNDANCES AND NUCLEOSYNTHESIS

    The ultimate goal of nucleosynthesis theory is, of course, to explain the composition of the Universe, as reflected, for example, in the stellar and gas components of galaxies. Significant progress has been achieved in this regard as a consequence of a wealth of new information of cosmic abundances—spectroscopic properties of stars in our galaxy and of gas clouds and galaxies at high redshifts—pouring in from new ground-and space-based observatories. Given that, it remains true that the most significant clues to nucleosynthesis are those provided by our detailed knowledge of the elemental and isotopic composition of solar system matter. The mass fractions of the stable isotopes in the solar are displayed in Figure 1. Key features that reflect the nature of the nuclear processes by which the heavy elements are formed include: (i) the large abundances of ¹²C and ¹⁶O, the main products of stellar helium burning; (ii) the dominance of the α-particle nuclei through calcium (²⁰Ne, ²⁴Mg, ²⁸Si, ³²S, ³⁶Ar, and ⁴⁰Ca); (iii) the nuclear statistical equilibrium peak at the position of ⁵⁶Fe; and (iv) the abundance peaks in the region past iron at the neutron closed shell positions (zirconium, barium, and lead), confirming the occurrence of processes of neutron-capture synthesis. The solar system abundance patterns associated specifically with the slow (s-process) and fast (r-process) processes of neutron capture synthesis are shown in Figure 2.

    Figure 1 The abundances of the isotopes present in solar system matter are plotted as a function of mass number A (the solar system abundances for the heavy elements are those compiled by Palme and Jones.

    Figure 2 The s-process and r-process abundances in solar system matter (based upon the work by Käppeler et al. , 1989 ). Note the distinctive s-process signature at masses A ~88, 138, and 208 and the corresponding r-process signatures at A ~130 and 195, all attributable to closed-shell effects on neutron capture cross-sections. It is the r-process pattern thus extracted from solar system abundances that can be compared with the observed heavy element patterns in extremely metal-deficient stars (the total solar system abundances for the heavy elements are those compiled by Anders and Grevesse, 1989 ), which are very similar to those from the compilation of Palme and Jones.

    It is important here to call attention to the revised determinations of the oxygen and carbon abundances in the Sun. Allende Prieto et al. (2001) derived an accurate oxygen abundance for the Sun of log ε(O) = 8.69 ± 0.05 dex, a value approximately a factor of 2 below that quoted by Anders and Grevesse (1989). Subsequently, Allende Prieto et al. (2002) determined the solar carbon abundance to be log ε(C) = 8.39 ± 0.04 dex, and the ratio C/O = 0.5 ± 0.07. The bottom line here is a reduction in the abundances of the two most abundant heavy elements in the Sun, relative to hydrogen and helium, by a factor ∼ 2. The implications of these results for stellar evolution, nucleosynthesis, the formation of carbon stars, and galactic chemical evolution remain to be explored.

    Guided by early compilations of the cosmic abundances as reflected in solar system material (e.g., Suess and Urey, 1956), Burbidge et al. (1957) and Cameron (1957) identified the nuclear processes by which element formation occurs in stellar and supernova environments: (i) hydrogen burning, which powers stars for ∼ 90% of their lifetimes; (ii) helium burning, which is responsible for the production of ¹²C and ¹⁶O, the two most abundant elements heavier than helium; (iii) the α-process, which we now understand as a combination of carbon, neon, and oxygen burning; (iv) the equilibrium process, by which silicon burning proceeds to the formation of a nuclear statistical equilibrium abundance peak centered on mass A = 56; (v) the slow (s-process) and rapid (r-process) mechanisms of neutron capture synthesis of the heaviest elements (A 60−70); and (vi) the p-process, a combination of the γ-process and the ν-process, which we understand to be responsible for the synthesis of a number of stable isotopes of nuclei on the proton-rich side of the valley of beta stability. Our subsequent discussions will identify the astrophysical environments in which these diverse processes are now understood to occur.

    1.3 INTERMEDIATE MASS STARS: EVOLUTION AND NUCLEOSYNTHESIS

    Intermediate-mass red giant stars are understood to be the primary source both of ¹²C and of the heavy s-process (slow neutron capture) elements, as well as a significant source of ¹⁴N and other less abundant CNO isotopes. Their contributions to galactic nucleosynthesis are a consequence of the occurrence of nuclear reactions in helium shell thermal pulses on the AGB, the subsequent dredge-up of matter into the hydrogen-rich envelope by convection, and mass loss. This is a very complicated evolution. Current stellar models, reviewed by Busso et al. (1999), allow the formation of low mass (~ 1.5M ) carbon stars, which represent the main source of s-process nuclei. A detailed review of the nucleosynthesis products (chemical yields) for low- and intermediate-mass stars is provided by Marigo (2001).

    Red giant stars have played a significant role in the historical development of nucleosynthesis theory. While the pivotal role played by nuclear reactions in stars in providing an energy source sufficient to power stars like the Sun over billions of years was established in the late 1930s, it remained to be demonstrated that nuclear processes in stellar interiors might play a role in the synthesis of heavy nuclei. The recognition that heavy-element synthesis is an ongoing process in stellar interiors followed the discovery by Merrill (1952) of the presence of the element technetium in red giant stars. Since technetium has no stable isotopes, and the longest-lived isotope has a half-life τ1/2 ∼ 4:6 Myr, its presence in red-giant atmospheres indicates its formation in these stars. This confirmed that the products of nuclear reactions operating at high temperatures and densities in the deep interior can be transported by convection to the outermost regions of the stellar envelope.

    The role of such convective dredge-up of matter in the red-giant phase of evolution of 1 −10M stars is now understood to be an extremely complex process (Busso et al., 1999). On the first ascent of the giant branch (prior to helium ignition), convection can bring the products of CNO cycle burning (e.g., ¹³C, ¹⁷O, and ¹⁴N) to the surface. A second dredge-up phase occurs following the termination of core helium burning. The critical third dredge up, occurring in the aftermath of thermal pulses in the helium shells of these AGB stars, is responsible for the transport of both ¹²C and s-process nuclei (e.g., technetium) to the surface. The subsequent loss of this enriched envelope matter by winds and planetary nebula formation serves to enrich the interstellar media of galaxies, from which new stars are born. A brief review of the mechanisms of production of ¹²C and the main component of the s-process of neutron capture nucleosynthesis is presented in the following sections.

    1.3.1 Shell Helium Burning and ¹²C Production

    Stellar helium burning proceeds by means of the triple-alpha reaction in which three ⁴He nuclei are converted into ¹²C, followed by the ¹²C(α, γ)¹⁶O reaction, which forms ¹⁶O at the expense of ¹²C. Core helium burning in massive stars (M 10M ) occurs at high temperatures, which increases the rate of the ¹²C(α, γ)¹⁶ O reaction and favors the production of oxygen. Typically, the ¹⁶O/¹²C ratio in the oxygen-rich mantles of massive stars prior to collapse is a factor ~ 2–3 higher than the solar values. Massive stars are thus the major source of oxygen, while low-and intermediate-mass stars dominate the production of carbon.

    The advantage of the helium shells of low- and intermediate-mass stars for ¹²C production arises from the fact that, for conditions of incomplete helium burning, the ¹²C/¹⁶O ratio is high. Following a thermal pulse in the helium shell, convective dredge-up of matter from the helium shell brings helium, s-process elements, and a significant mass of ¹²C to the surface. It is the surface enrichment associated with this source of ¹²C that leads to the condition that the envelope ¹²C/¹⁶O ratio exceeds 1, such that carbon star is born. Calculations of galactic chemical evolution indicate that this source of carbon is sufficient to account for the level of ¹²C in galactic matter. The levels of production of ¹⁴N, ¹³C, and other CNO isotopes in this environment are significantly more difficult to estimate, and thus the corresponding contributions of AGB stars to the galactic abundances of these isotopes remain uncertain.

    1.3.2 s-Process Synthesis in Red Giants

    The formation of most of the heavy elements occurs in one of two processes of neutron capture: the s-process or the r-process. These two broad divisions are distinguished on the basis of the relative lifetimes for neutron captures (τn) and electron decays (τβ). The condition that τn > τβ, where τβ is a characteristic lifetime for β-unstable nuclei near the valley of β-stability, ensures that as captures proceed the neutron-capture path will itself remain close to the valley of β-stability. This defines the s-process. In contrast, when τn < τβ, it follows that successive neutron captures will proceed into the neutron-rich regions off the β-stable valley. Following the exhaustion of the neutron flux, the capture products approach the position of the valley of β-stability by β-decay, forming the r-process nuclei. The s-process and r-process patterns in solar system matter are those shown in Figure 2.

    The environment provided by thermal pulses in the helium shells of intermediate-mass stars on the AGB provides conditions consistent with the synthesis of the bulk of the heavy s-process isotopes through bismuth. Neutron captures in AGB stars are driven by a combination of neutron sources: the ¹³C(α, n)¹⁶O reaction provides the bulk of the neutron budget at low-neutron densities, while the ²²Ne(α, n)²⁵ Mg operating at high temperatures helps to set the timescale for critical reaction branches. This s-process site (the main s-process component) is understood to operate in low-mass AGB stars (M~ 1 − 3M ) and to be responsible for the synthesis of the s-process nuclei in the mass range A 90. Calculations reviewed by Busso et al. (1999) indicate a great sensitivity both to the characteristics of the ¹³C pocket in which neutron production occurs and to the initial metalli-city of the star. In their view, this implies that the solar system abundances are not the result of a unique s-process but rather the consequence of a complicated galactic chemical evolutionary history which witnessed mixing of the products of s-processing in stars of different metallicity and a range of ¹³C pockets. We can hope that observations of the s-process abundance patterns in stars as a function of metallicity will ultimately be better able to guide and to constrain such theoretical models.

    1.4 MASSIVE-STAR EVOLUTION AND NUCLEOSYNTHESIS

    Generally speaking, the evolution of a massive star follows a well-understood path of contraction to increasing central density and temperature. The contraction is caused by the energy loss of the star, due to light radiated from the surface and neutrino losses (see below). The released potential energy is in part converted into internal energy of the gas (Virial theorem). This path of contraction is interrupted by nuclear fusion—first hydrogen is burned to helium, then helium to carbon and oxygen. This is followed by stages of carbon, neon, oxygen, and silicon burning, until finally a core of iron is produced, from which no more energy can be extracted by nuclear burning. The onsets of these burning phases as the star evolves through the temperature–density plane are shown in Figure 3, for stars of masses 15M and 25M . Each fuel burns first in the center of the star, then in one or more shells (Figure 4). Most burning stages proceed convectively: i.e., the energy production rate by the burning is so large and centrally concentrated that the energy cannot be transported by radiation (heat diffusion) alone, and convective motions dominate the heat transport. The reason for this is the high-temperature sensitivity of nuclear reaction rates: for hydrogen burning in massive stars, nuclear energy generation has a ∝ T¹⁸ dependence, and the dependence is even stronger for later burning stages. The important consequence is that, due to the efficient mixing caused by the convection, the entire unstable region evolves essentially chemically homogeneously—replenishing the fuel at the bottom of burning region (central or shell burning) and depleting the fuel elsewhere in that region at the same time. As a result, shell burning of this fuel then commences outside that region.

    Figure 3 Evolution of the central temperature and density in stars of 15 M and 25 M from birth as hydrogen burning stars until iron core collapse ( Table 1 ). In general, the trajectories follow a line of ρ T ³ , but with some deviation downward (towards higher ρ at a given T ) due to the decreasing entropy of the core. Nonmonotonic behavior is observed when nuclear fuels are ignited and this is exacerbated in the 15 M model by partial degeneracy of the gas (source Woosley et al. , 2002 ).

    Figure 4 Interior structure of a 22M star of solar composition as a function of time (logarithm of time till core collapse) and enclosed mass. Green hatching and red cross hatching indicate convective and semiconvective regions. Convective regions are typically well mixed and evolve chemically homogeneously. Blue shading indicates energy generation and pink shading energy loss. Both take into account the sum of nuclear and neutrino loss contributions. The thick black line at the top indicates the total mass of the star, being reduced by mass loss due to stellar winds. Note that the mass loss rate actually increases at late times of the stellar evolution. The decreasing slope of the total mass of the star in the figure is due to the logarithmic scale chosen for the time axis.

    Table 1 summarizes the burning stages and their durations for a 20M star. The timescale for helium burning is ∼ 10 times shorter than that of hydrogen burning, mostly because of the lower energy release per unit mass. The timescale of the burning stages and contraction beyond central helium burning is greatly reduced by thermal neutrino losses that carry away energy in situ, instead of requiring that it be transported to the stellar surface by diffusion or convection. These losses increase with temperature (as ∝ T⁹). When the star has built up a large-enough iron core, exceeding its effective Chandrasekhar mass (the maximum mass for which such a core can be stable), the core collapses to form a neutron star or a black hole (see Woosley et al. (2002) for a more extended review). A supernova explosion may result (e.g., Colgate and White, 1966) that ejects most of the layers outside the iron core, including many of the ashes from the preceding burning phase. However, when the supernova shock front travels outward, for a brief time peak temperatures are reached, that exceed the maximum temperatures that have been reached in each region in the preceding hydrostatic burning stages (Table 2). This defines the transient stage of explosive nucleosynthesis that is critical to the formation of an equilibrium peak dominated by ⁵⁶Ni (Truran et al., 1967).

    Table 1

    Hydrostatic nuclear burning stages in massive stars. The table gives burning stages, main and secondary products (ashes), typical temperatures and burning timescales for a 20M star, and the main nuclear reactions. An ellipsis (…) indicates more than one product of the double carbon and double oxygen reactions, and a chain of reactions leading to the buildup of iron group elements for silicon burning.

    Table 2

    Explosive nucleosynthesis in supernovae. Similar to Table 1, but for the explosion of the star. The (A, B) notation means A is on the ingoing channel and B is on the outgoing channel. An "α is same as He, γ denotes a photon, i.e., a photodisintegration reaction when on the ingoing channel, n" is a neutron, β− shows β-decay, and ν indicates neutrino-induced reactions.

    Massive stars build up most of the heavy elements from oxygen through the iron group from the initial hydrogen and helium of which they are formed. They also make most of the s-process heavy elements up to atomic mass numbers 80–90 from initial iron, converting initial carbon, oxygen, and nitrogen into ²²Ne, thus providing a neutron source for the s-process. Massive stars are probably also the source of most of the proton-rich isotopes of atomic mass number greater than 100 (p-process) and the site of the r-process that is responsible for many of the neutron-rich isotopes from barium through uranium and thorium.

    1.4.1 Nucleosynthesis in Massive Stars

    The most abundant product of the evolution of massive stars is oxygen, ¹⁶O in particular—the third most abundant isotope in the Universe and the most abundant metal. Massive stars are also the main source of most heavy elements up to atomic mass number A ~80, of some of the rare proton-rich nuclei, and of the r-process nuclei from barium to uranium. In the following, we will briefly review the burning stages and nuclear processes that characterize the evolution of massive stars and the resulting core collapse supernovae.

    1.4.1.1 Hydrogen burning

    The first stage of stellar burning and energy generation is hydrogen burning, during which hydrogen is transformed to helium by the so-called CNO cycle. An initial abundance of carbon, nitrogen, and oxygen (of which ¹⁶O is the most abundant) is collected into ¹⁴N—the proton capture on this isotope is the slowest reaction in this cycle. The total number of CNO isotopes remains unchanged—they operate only as nuclear catalysts in the conversion of hydrogen into helium.

    1.4.1.2 Helium burning and the s-process

    After hydrogen is depleted, the star contracts towards helium burning. Before helium burning is ignited, the ¹⁴N is converted into ¹⁸O by ¹⁴N(α, γ)¹⁸F(β+)¹⁸ O and then burned to ²²Ne by ¹⁸O(α, γ)²²Ne, where α stands for a ⁴He nucleus and the bracket notation means that the first particle is on the ingoing channel and the second on the outgoing channel. The first stage of helium burning involves the interaction of three α-particles to form carbon (⁴He(α, γ)⁸ Be*(α, γ) ¹²C; "3α reaction"). As helium becomes depleted, the a-capture on ¹²C takes over and produces ¹⁶O. Indeed, as helium is entirely depleted in the central regions of the star, most of the ashes are ¹⁶O, while ¹²C only comprises 10–20%, decreasing with increasing stellar mass.

    Towards the end of the central helium burning phase, the temperature becomes high enough for the ²²Ne(α, n) ²⁵Mg reaction to proceed, which provides a neutron source for the so-called weak component of the s-process. In massive stars, this process builds up elements of atomic mass number of up to 80–90, starting with neutron capture on original ⁵⁶Fe present in the interstellar medium from which the star was born. In this environment, the timescale for neutron capture is slow compared to the weak (β−) decay of radioactive nuclei produced by the capture. The s-process path thus proceeds along the neutron-rich side of the valley of stable nuclei. The main competitor reaction for the neutron source is the destruction channel for ²²Ne that does not produce neutrons, ²²Ne(α,γ)²⁶Mg. The main neutron sink or poison for the s-process in this environment is neutron captures on the progeny of the ²²Ne(α, n) reaction, among which ²⁵Mg is itself a major neutron poison.

    1.4.1.3 Hydrogen and helium shell burning

    During central helium burning, hydrogen continues to burn in a shell at the outer edge of the helium core, leaving the ashes (helium) at the base of the shell and increasing the size of the helium core. This growth of the helium core by the continued addition of fresh helium fuel contributes to the destruction of carbon at the end of helium burning.

    At the completion of core helium burning, helium continues to burn in a shell overlying the helium-free core. It first burns radiatively but then becomes convective, especially in more massive stars. Since helium is not completely consumed until the death of the star, some s-processing may start here as well, as a consequence of the higher temperatures characterizing the shell burning phase. By the time the stable burning phases of evolution are completed and the iron core is on the verge of collapse, this shell consists of a mixture of helium, carbon, and oxygen, each of which comprises several tens of percent of the matter.

    1.4.1.4 Carbon burning

    After a contraction phase of several 10⁴ yr carbon burning starts in the center. This produces mostly ²⁰Ne(via ¹²C(¹²C, α)²⁰Ne), but also some ²⁴Mg(via ¹²C(¹²C, γ)²⁴Mg) and ²³Na(via ¹²C(¹²C, p)²³Na). There is also some continuation of the s-process from unburned ²²Ne after the end of central helium burning, operating with neutrons released during carbon burning via the reaction ¹²C(¹²C, n)²³Mg.

    1.4.1.5 Neon and oxygen burning

    Neon burning is induced by photodisintegration of ²⁰Ne into an α-particle and ¹⁶O. The α-particle is then captured by another ²⁰Ne nucleus to make ²⁴Mg or by ²⁴Mg to make ²⁸Si. Secondary products include ²⁷Al, ³¹P, and ³²S.

    This is closely followed by oxygen burning, for which the dominant reaction is ¹⁶O + ¹⁶O. The significant products are the α-particle nuclei ²⁸Si, ³²S, ³⁶Ar, and ⁴⁰Ca, together with such isotopes of odd-Z nuclei such as ³⁵Cl, ³⁷Cl, and ³⁹Cl.

    1.4.1.6 Silicon burning

    Finally, ²⁸Si (and ³²S) is burned in a similar way as ²⁰Ne: photodisintegration of ²⁸Si and resulting lighter isotopes accompanies the gradual buildup of iron peak nuclei with higher binding energies per nucleon. At the same time weak processes—positron decaying and electron capture—also become important, producing an neutron excess and allowing the formation of nuclei along the valley of stability. The ultimate result is the formation of an iron equilibrium peak of nuclei of maximal nuclear binding energies, from which no further energy can be gained by means of nuclear fusion. The extent in mass of convective central silicon burning is typically ~1.05M c. Several brief stages of shell silicon burning follow until the star has built a core of iron group elements (iron core) large enough to collapse, typically 1.3M 2M . The core had been held up, in part, by electron degeneracy pressure, but when their Fermi energy becomes sufficiently large they can be captured on protons. This serves both to neutronize the core and to reduce pressure support, ultimately leading to the collapse of the core to a neutron star or black hole. It follows that virtually all of the iron-peak products of this phase of quasi-hydrostatic silicon burning do not escape from the star and thus do not contribute to galactic nucleosynthesis. It is the overlying regions of the core that will generally be ejected.

    1.4.1.7 Explosive nucleosynthesis

    Following the collapse of the core a supernova shock front, driven by energy deposition by neutrinos from the hot protoneutron star, travels outwards through the star, setting the stage for a brief phase of explosive nucleosynthesis. In the inner regions, in silicon- and oxygen-rich layers, the photodisintegration of silicon into a-particles and free nucleons is accompanied by the capture of these particles onto silicon and heavier nuclei, leading to the formation of a nuclear statistical equilibrium peak centered on mass A = 56. Since the material here is characterized by a small excess of neutrons over protons, it follows that the final products of these explosive burning episodes must lie along or very near to the Z = N line. The dominant species in situ therefore include the nuclei ⁴⁴Ti, ⁴⁸Cr, ⁵²Fe, ⁵⁶Ni, and ⁶⁰Zn. Following decay these contribute to the production of the most proton-rich stable isotopes at these mass numbers: ⁴⁴Ca, ⁴⁸Ti, ⁵²Cr, ⁵⁶Fe, and ⁶⁰Ni. The neutron enrichments characteristic of matter of solar composition also allow the production of the odd-Z species ⁵¹V (formed as ⁵¹Mn), ⁵⁵Mn (formed as ⁵⁵Co), and ⁵⁹Co (formed as ⁵⁹Cu).

    Further out, in the oxygen, neon, and carbon shells, explosive burning can act to modify somewhat the abundance patterns resulting from earlier hydrostatic burning phases. The final nucleosynthesis products of the evolution of massive stars and associated type II core collapse supernovae have been discussed by a number of authors (Woosley and Weaver, 1995; Thielemann et al., 1996; Nomoto et al., 1997; Limongi et al., 2000; Woosley et al., 2002). Most of the elements and isotopes in the region from ¹⁶O to ⁴⁰Ca are formed in relative proportions consistent with solar system matter. The same may be said of the nuclei in the iron-peak region from approximately ⁴⁸Ti to ⁶⁴Zn. A unique signature of such massive star/type II supernova nucleosynthesis, however, is the fact that nuclei in the ¹⁶O to ⁴⁰Ca region are overproduced by a factor ~ 2–3 with respect to nuclei in the ⁴⁸Ti to ⁶⁴Zn region—relative to solar system abundances. We will see how this signature, reflected in the compositions of the oldest stars in our galaxy provides important constraints on models of galactic chemical evolution.

    1.4.1.8 The p-process

    The p-process generally serves to include all possible mechanisms that can contribute to the formation of proton-rich isotopes of heavy nuclei. It was previously believed that these nuclei might be formed by proton captures. We now understand that, for the conditions that obtain in massive stars and accompanying supernovae, proton captures are not sufficient. The two dominant processes making proton-rich nuclei in massive stars are the γ-process and the ν-process.

    The γ-process involves the photodisintegration of heavy elements. Obviously, this process is not very efficient, as the effective seed nuclei are the primordial heavy-element constituents of the star. Most of the heavy p-elements with atomic mass number greater than 100 that are produced in massive stars (see Figure 5) are formed by this mechanism. It is primarily for this reason that these proton-rich heavy isotopes are rare in nature.

    Figure 5 Production factor in the ejecta of a 15M star from Rauscher et al. (2002). The production factor is defined as the average abundance of the isotope in the stellar ejecta divided by its solar abundance. To guide the eye, we mark the production factor of ¹⁶O by a dashed line and a range of acceptable co-production, i.e., within a factor 2 of ¹⁶O, by dotted lines. The absolute value of the production factors does not matter so much, as the ejecta will be diluted with the interstellar material after the supernova explosion. However, to reproduce a solar abundance pattern for certain isotopes in massive stars, they should be closely co-produced with the most abundant product of massive stars, ¹⁶O.

    The rarest of these proton-rich isotopes have a different and more exotic sources: the immense neutrino flux from the forming hot neutron star can either convert a neutron into a proton (making, e.g., ¹³⁸La, ¹⁸⁰Ta) or knock out a nucleon (e.g., ¹¹B, ¹⁹F). Since the neutrino cross-sections are quite small, it is clear that the parent nucleus must have an abundance that is at least several thousand times greater than that of the neutrino-produced daughter if this process is to contribute significantly to nucleosynthesis. This is indeed true for the cases of ¹¹B (parent ¹²C) and ¹⁹F (parent ²⁰Ne). Production of La and ¹⁸⁰Ta, the two rarest stable isotopes in the Universe, is provided by neutrino interactions with ¹³⁸Ba and ¹⁸⁰Hf, respectively.

    1.4.1.9 The r-process

    It is generally accepted that the r-process synthesis of the heavy neutron capture elements in the mass regime A 130—140 occurs in an environment associated with massive stars. This results from two factors: (i) the two most promising mechanisms for r-process synthesis—a neutrino heated hot bubble and neutron star mergers— are both tied to environments associated with core collapse supernovae; and (ii) observations of old stars (discussed in Section 1.6) confirm the early entry of r-process isotopes into galactic matter.

    • The r-process model that has received the greatest study in recent years involves a high-entropy (neutrino-driven) wind from a core collapse supernova (Woosley et al., 1994; Takahashi et al., 1994). The attractive features of this model include the facts that it may be a natural consequence of the neutrino emission that must accompany core collapse in collapse events, that it would appear to be quite robust, and that it is indeed associated with massive stars of short lifetime. Recent calculations have, however, called attention to a significant problem associated with this mechanism: the entropy values predicted by current type II supernova models are too low to yield the correct levels of production of both the lighter and heavier r-process nuclei.

    • The conditions estimated to characterize the decompressed ejecta from neutron star mergers (Lattimer et al., 1977; Rosswog et al., 1999) may also be compatible with the production of an r-process abundance pattern generally consistent with solar system matter. The most recent numerical study of r-process nucleosynthesis in matter ejected in such mergers (Freiburghaus et al., 1999) show specifically that the r-process heavy nuclei in the mass range A 130—140 are produced in solar proportions. Here again, the association with a massive star/core collapse super-nova environment is consistent with the early appearance of r-process nuclei in the galaxy and the mechanism seems quite robust.

    Massive stars may also contribute to the abundances of the lighter r-process isotopes (A 130—140). The helium and carbon shells of massive stars undergoing supernova explosions can give rise to neutron production via such reactions as ¹³C(α, n)¹⁶O, ¹⁸O(α, n)²¹Ne, and ²²Ne(α, n)²⁵Mg, involving residues of hydrostatic burning phases. Early studies of this problem (Hillebrandt et al., 1981; Truran et al., 1978) indicated that these conditions might allow the production of at least the light (A 130—140) r-process isotopes (see also Truran and Cowan (2000)). A recent analysis of nucleosynthesis in massive stars by Rauscher et al. (2002) has, however, found that this yields only a slight redistribution of heavy-mass nuclei at the base of the helium shell and no significant production of r-process nuclei above mass A = 100. This environment may, nevertheless, provide a source of r-process-like anomalies in grains.

    1.5 TYPE Ia SUPERNOVAE: PROGENITORS AND NUCLEOSYNTHESIS

    Two broad classes of supernovae are observed to occur in the Universe: type I and type II. We have learned from our discussion in the previous section that type II (core collapse) supernovae are products of the evolution of massive stars (M 10M ). Observationally, the critical distinguishing feature of type I supernovae is the absence of hydrogen features in their spectra at maximum light. Theoretical studies focus attention on models for type I events involving either exploding white dwarfs (type Ia) or the explosions of massive stars (similar to those discussed previously) which have, via wind-driven mass loss or binary effects, shed virtually their entire hydrogen envelopes prior to (iron) core collapse (e.g., types Ib and Ic). We are concerned here with supernovae of type Ia, a subclass of the type Is, which are understood to be associated with the explosion of a white dwarf in a close binary system. The standard model for SNe Ia involves specifically the growth of a carbon–oxygen white dwarf to the Chandrasekhar-limit mass in a close binary system and its subsequent incineration (see, e.g., the review of type Ia progenitor models by Livio, 2000).

    The iron-peak nuclei observed in nature had their origin in supernova explosions. Type II and type Ia supernovae provide the dominant sites in which this explosive nucleosynthesis mechanism is known to operate. These two supernova sites operate on distinctively different timescales and eject different amounts of iron. As we shall see, an understanding of the detailed nucleosynthesis patterns emerging from these two classes of events provides important insights into the star formation histories of galaxies. Type II (core collapse) supernovae produce both the intermediate-mass nuclei from oxygen to calcium-and iron-peak nuclei. Calculations of charged-particle nucleosynthesis in massive stars and type II supernovae (Woosley and Weaver, 1995; Thielemann et al., 1996; Nomoto et al., 1997; Limongi et al., 2000) reveal one particularly significant distinguishing feature of the emerging abundance patterns: the elements from oxygen through calcium (and titanium) are overproduced relative to iron (peak nuclei) by a factor ~ 2–3. This means that SNe II produce only ~1/3–1/2 of the iron in galactic matter. (We note that, while all such models necessarily make use of an artificially induced shock wave via thermal energy deposition (Thielemann et al., 1996) or a piston (Woosley and Weaver, 1995), the general trends obtained from such nucleosynthesis studies are expected to be valid.) As we shall see in our discussion of chemical evolution, these trends are reflected in the abundance patterns of metal deficient stars in the halo of our galaxy (Wheeler et al., 1989; McWilliam, 1997). This leaves to SNe Ia the need to produce the ~1/3–1/2 of the iron-peak nuclei from titanium to zinc (Ti–V–Cr–Mn–Fe–Co–Ni–Cu–Zn).

    Calculations of explosive nucleosynthesis associated with carbon deflagration models for type Ia events (Thielemann et al., 1986; Iwamoto et al., 1999; Dominguez et al., 2001) predict that sufficient iron-peak nuclei ∼(0.6−0.8M of ⁵⁶Fe in the form of ⁵⁶Ni) are synthesized to explain both the powering of the light curves due to the decays of ⁵⁶Ni and ⁵⁶Co and the observed mass fraction of ⁵⁶Fe in galactic matter. Estimates of the timescale for first entry of the ejecta of SNe Ia into the interstellar medium of our galaxy yield ∼ 2 × 10⁹ yr. at a metallicity [Fe/H]~− 1 (Kobayashi et al., 2000; Goswami and Prantzos, 2000).

    A representative nucleosynthesis calculation for such a type Ia supernova event is shown in Figure 6. Note particularly the region of mass fraction between ∼ 0.2M and 0.8M , which is dominated by the presence of ⁵⁶Ni, is in nuclear statistical equilibrium. It is this nickel mass that is responsible—as a consequence of the decay of ⁵⁶Ni through ⁵⁶Co to ⁵⁶Fe—for the bulk of the luminosity of type Ia supernovae at maximum light. This is a critical factor in making SNe Ia the tool of choice for the determination of the cosmological distance scale—as they are the brightest stellar objects known. From the point of view of nucleosynthesis, we then understand that it is the ~ 0.6M of iron-peak nuclei ejected per event by type Ia supernovae that represents the bulk of ⁵⁶Fe in galactic matter. The remaining mass, which is converted into nuclei from ¹⁶O to ⁴⁰Ca, does not make a significant contribution to the synthesis of these elements.

    Figure 6 within which the dominant product is ⁵⁶ Ni. (source Timmes et al., 2003)

    1.6

    NUCLEOSYNTHESIS AND GALACTIC CHEMICAL EVOLUTION

    M = M 10 of intermediate-mass stars, for which substantial element production occurs during the AGB phase of their evolution; (ii) the mass range M 10M , corresponding to the massive star progenitors of type II (core collapse) supernovae; and (iii) type Ia supernovae, which are understood to arise as a consequence of the evolution of intermediate mass stars in close binary systems.

    In the context of models of galactic chemical evolution, it is extremely important to know as well the production timescales for each of these sites—i.e., the effective timescales for the return of a star’s nucleosynthesis yields to the interstellar gas. The lifetime of a 10M star is ∼ 5 × 10⁷yr. We can thus expect all massive stars M 10M to evolve on timescales τSNII < 10⁸ yr, and to represent the first sources of heavy-element enrichment of stellar populations. In contrast, intermediate-mass stars evolve on timescales τ10⁸ − 10⁹yr. Finally the timescale for SNe Ia product enrichment is a complicated function of the binary history of type Ia progenitors (see, e.g., Livio, 2000). Observations and theory suggest a timescale τ(1.5−2) × 10⁹yr.

    Many significant features of the evolution of our galaxy follow from a knowledge of the nuclear ashes and evolutionary timescales for the three sites we have surveyed. The primordial composition of the galaxy was that which it inherited from cosmological nucleosynthesis—primarily hydrogen and helium. The characteristics of the first stellar contributions to nucleosynthesis, whether associated with population II or population III, reflect (as might be expected) the nucleosynthesis products of the evolution of massive stars (M 10M ) of short lifetimes (τ 10⁸yr). The significant trends in galactic chemical evolution of concen here are those involving the timing of first entry of the products of the other two broad classifications of nucleosynthesis contributors that we have identified: low-mass stars (s-process) and type Ia supernovae (iron-peak nuclei).

    Spectroscopic abundance studies of the oldest stars in our galaxy, down to metallicities [Fe/H] ~−4 to −3, have been reviewed by Wheeler et al. − 3) in our own galaxy’s halo show two significant variations with respect to solar abundances: the elements in the mass range from oxygen to calcium are overabundant—relative to iron-peak nuclei—by a factor ~ 2–3, and the abundance pattern in the heavy element region A 60−70 closely reflects the r-process abundance distribution that is characteristic of solar system matter, with no evidence for an s-process nucleosynthesis contribution. For purposes of illustration, the trends in [Ca/Fe] as a function of metallicity [Fe/H] are shown in Figure 7. Note the factor 2–3 overabundance of calcium with respect to iron at metallicities below [Fe/H] ~− 1. We also display in Figure 8 the detailed agreement of metal-poor star heavy-element abundance pattern with that of solar system r-process abundances, for three metal-poor but r-process-rich stars: CS 22892-052 (Sneden et al., 2003), HD 115444 (Westin et al., 2000), and BD+17°3248 (Cowan et al., 2002). Both of these features are entirely consistent with nucleosynthesis expectations for massive stars and associated type II supernovae.

    Figure 7 : : Edvardsson et al. , (1993) ).

    Figure 8 The heavy-element abundance patterns for the three stars CS 22892-052, HD 155444, and BD+17°3248 are compared with the scaled solar system r-process abundance distribution (solid line) (Sneden et al., 2003; Westin et al., 2000; Cowan et al., 2002). Upper limits are indicated by inverted triangles (source Truran et al., 2003).

    The signatures of an increasing s-process contamination first appears at an [Fe/H] ~− 2.5 to − 2.0. The evidence for this is provided by an increase in the ratio of barium (the abundance of which in galactic matter is dominated by s-process contributions) to europium (almost exclusively an r-process product). The ratio [Ba/Eu] is shown in Figure 9 as a function of [Fe/H] for a large sample of halo and disk stars. Note that at the lowest metallicities the [Ba/Eu] ratio clusters around the pure r-process value ([Ba/Eu]r-process~– 0.9); at a metallicity [Fe/H] ~–2.5, the Ba/Eu ratio shows a gradual increase. In the context of our earlier review of nucleosynthesis sites, this provides observational evidence for the first input from AGB stars, on timescales perhaps approaching τ10⁹ yr.

    Figure 9 : Burris et al. (2000) ; ×: Woolf et al. (1995) and Edvardsson et al. : Zhao and Magain (1990) ).

    In contrast, evidence for entry of the iron-rich ejecta of SNe Ia is seen first to appear at a metallicity [Fe/H]~− 1.5 to − 1.0. This may be seen reflected in the abundance histories of [Ca/Fe] with [Fe/H], in Figure 7. This implies input from supernovae Ia on timescales τ(1 − 2) × 10⁹ yr. It may be of interest that this seems to appear at approximately the transition from halo to (thick) disk stars.

    These observed abundance histories confirm theoretical expectations for the nucleosynthesis contributions from different stellar and supernova sites and the timescales on which enrichment occurs. They provide extremely important tests of numerical simulations and illustrate how the interplay of theory and observation can provide important constraints on the star formation and nucleosynthesis history of our galaxy.

    REFERENCES

    Allende Prieto C., Lambert DL, Asplund M. The forbidden abundance of oxygen in the Sun. Astrophys. J. 2001;556:L63–L66.

    Allende Prieto C., Lambert DL, Asplund M. A reappraisal of the solar photospheric C/O ratio. Astrophys. J. 2002;573:L137–L140.

    Anders E, Grevesse N. Abundances of the elements: meteoritic and solar. Geochim. Cosmochim. Acta. 1989;53:197–214.

    Arnett WD. Explosive nucleosynthesis revisited: yields. Ann. Rev. Astron. Astrophys. 1995;33:115–132.

    Burbidge EM, Burbidge GR, Fowler WA, Hoyle F. Synthesis of the elements in stars. Rev. Mod. Phys. 1957;29:547–650.

    Burles S, Nollett KM, Turner MS. Big Bang nucleosynthesis predictions for precision cosmology Astrophys. J. 2001;552:L1–L6.

    Burris DL, Pilachowski CA, Armandroff TA, Sneden C, Cowan JJ, Roe H. Neutron-capture elements in the early galaxy: insights from a large sample of metal-poor giants. Astrophys. J. 2000;544:302–319.

    Busso M, Gallino R, Wasserburg GJ. Nucleosynthesis in asymptotic giant branch stars: relevance for galactic enrichment and solar system formation. Ann. Rev Astron. Astrophys. 1999;37:239–309.

    Cameron AGW. Stellar Evolution Nuclear Astrophysics and Nucleogenesis. Chalk River Report, AELC (Atomic Energy Canada). 1957 CRL-41.

    Colgate SA, White RH. The hydrodynamic behavior of supernovae explosions. Astrophys. J. 1966;143:626–681.

    Cowan JJ, Sneden C, Burles S, Ivans II, Beers TC, Truran JW, Lawler JE, Primas F, Fuller GM, Pfeiffer B, Kratz K-L. The chemical composition and age of the metal-poor halo star BD + 17°3248. Astrophys. J. 2002;572:861–879.

    Cyburt RH, Fields BD, Olive KA. Primordial nucleosynthesis with CMB inputs: probing the early universe and light element astrophysics. Astropart. Phys. 2002;17:87–100.

    Dominguez I, Höflich P, Straniero O. Constraints on the progenitors of Type Ia supernovae and implications for the cosmological equation of state. Astrophys. J. 2001;557:279–291.

    Edvardsson B, Andersen J, Gustafsson B, Lambert DL, Nissen PE, Tomkin J. The chemical evolution of the galactic disk I. Analysis and results. Astron. Astrophys. 1993;275:101–152.

    Freiburghaus C, Rosswog S, Thielemann F-K. r-Process in neutron star mergers. Astrophys. J. 1999;525:L121–L124.

    Gehrz RD, Truran JW, Williams RE, Starrfield S. Nucleosynthesis in classical novae and its contribution to the interstellar medium. Proc. Astron. Soc. Pacific. 1998;110:3–26.

    Goswami A, Prantzos N. Abundance evolution of intermediate mass

    Enjoying the preview?
    Page 1 of 1