Stellar nucleosynthesis

Source: Wikipedia, the free encyclopedia.
Logarithmic scale plot of the relative energy output (ε) of the following fusion processes at different temperatures (T):
  Combined energy generation of PP and CNO within a star
  The Sun's core temperature, at which PP is more efficient

In astrophysics, stellar nucleosynthesis is the creation of chemical elements by nuclear fusion reactions within stars. Stellar nucleosynthesis has occurred since the original creation of hydrogen, helium and lithium during the Big Bang. As a predictive theory, it yields accurate estimates of the observed abundances of the elements. It explains why the observed abundances of elements change over time and why some elements and their isotopes are much more abundant than others. The theory was initially proposed by Fred Hoyle in 1946,[1] who later refined it in 1954.[2] Further advances were made, especially to nucleosynthesis by neutron capture of the elements heavier than iron, by Margaret and Geoffrey Burbidge, William Alfred Fowler and Fred Hoyle in their famous 1957 B2FH paper,[3] which became one of the most heavily cited papers in astrophysics history.

helium (horizontal branch star), and progressively burning higher elements. However, this does not by itself significantly alter the abundances of elements in the universe as the elements are contained within the star. Later in its life, a low-mass star will slowly eject its atmosphere via stellar wind, forming a planetary nebula, while a higher–mass star will eject mass via a sudden catastrophic event called a supernova. The term supernova nucleosynthesis is used to describe the creation of elements during the explosion of a massive star or white dwarf
.

The advanced sequence of burning fuels is driven by

silicon. However, most of the nucleosynthesis in the mass range A = 28–56 (from silicon to nickel) is actually caused by the upper layers of the star collapsing onto the core, creating a compressional shock wave rebounding outward. The shock front briefly raises temperatures by roughly 50%, thereby causing furious burning for about a second. This final burning in massive stars, called explosive nucleosynthesis or supernova nucleosynthesis
, is the final epoch of stellar nucleosynthesis.

A stimulus to the development of the theory of nucleosynthesis was the discovery of variations in the

history of nucleosynthesis theory).[4] This suggested a natural process that is not random. A second stimulus to understanding the processes of stellar nucleosynthesis occurred during the 20th century, when it was realized that the energy released from nuclear fusion reactions accounted for the longevity of the Sun as a source of heat and light.[5]

History

In 1920, Arthur Eddington proposed that stars obtained their energy from nuclear fusion of hydrogen to form helium and also raised the possibility that the heavier elements are produced in stars.

In 1920,

strong nuclear force which is effective only at very short distances.[9]: 410  In the following decade the Gamow factor was used by Atkinson and Houtermans and later by Edward Teller
and Gamow himself to derive the rate at which nuclear reactions would occur at the high temperatures believed to exist in stellar interiors.

In 1939, in a

proton–proton chain reaction, is the dominant energy source in stars with masses up to about the mass of the Sun. The second process, the carbon–nitrogen–oxygen cycle, which was also considered by Carl Friedrich von Weizsäcker in 1938, is more important in more massive main-sequence stars.[11]: 167  These works concerned the energy generation capable of keeping stars hot. A clear physical description of the proton–proton chain and of the CNO cycle appears in a 1968 textbook.[12]: 365  Bethe's two papers did not address the creation of heavier nuclei, however. That theory was begun by Fred Hoyle in 1946 with his argument that a collection of very hot nuclei would assemble thermodynamically into iron.[1] Hoyle followed that in 1954 with a paper describing how advanced fusion stages within massive stars would synthesize the elements from carbon to iron in mass.[2][13]

Hoyle's theory was extended to other processes, beginning with the publication of the 1957 review paper "Synthesis of the Elements in Stars" by Burbidge, Burbidge, Fowler and Hoyle, more commonly referred to as the B2FH paper.[3] This review paper collected and refined earlier research into a heavily cited picture that gave promise of accounting for the observed relative abundances of the elements; but it did not itself enlarge Hoyle's 1954 picture for the origin of primary nuclei as much as many assumed, except in the understanding of nucleosynthesis of those elements heavier than iron by neutron capture. Significant improvements were made by Alastair G. W. Cameron and by Donald D. Clayton. In 1957 Cameron presented his own independent approach to nucleosynthesis,[14] informed by Hoyle's example, and introduced computers into time-dependent calculations of evolution of nuclear systems. Clayton calculated the first time-dependent models of the s-process in 1961[15] and of the r-process in 1965,[16] as well as of the burning of silicon into the abundant alpha-particle nuclei and iron-group elements in 1968,[17][18] and discovered radiogenic chronologies[19] for determining the age of the elements.

Cross section of a supergiant showing nucleosynthesis and elements formed.

Key reactions

A version of the periodic table indicating the origins – including stellar nucleosynthesis – of the elements.

The most important reactions in stellar nucleosynthesis:

Hydrogen fusion

Proton–proton chain reaction
CNO-I cycle
The helium nucleus is released at the top-left step.

Hydrogen fusion (nuclear fusion of four protons to form a

white dwarfs, are fusing hydrogen by these two processes.[21]
: 245 

In the cores of lower-mass main-sequence stars such as the

As a result, there is little mixing of fresh hydrogen into the core or fusion products outward.

In higher-mass stars, the dominant energy production process is the

convective energy transfer becomes more important than does radiative transfer. As a result, the core region becomes a convection zone, which stirs the hydrogen fusion region and keeps it well mixed with the surrounding proton-rich region.[26] This core convection occurs in stars where the CNO cycle contributes more than 20% of the total energy. As the star ages and the core temperature increases, the region occupied by the convection zone slowly shrinks from 20% of the mass down to the inner 8% of the mass.[25] The Sun produces on the order of 1% of its energy from the CNO cycle.[27][a][28]: 357 [29][b]

The type of hydrogen fusion process that dominates in a star is determined by the temperature dependency differences between the two reactions. The proton–proton chain reaction starts at temperatures about 4×106 K,[30] making it the dominant fusion mechanism in smaller stars. A self-maintaining CNO chain requires a higher temperature of approximately 1.6×107 K, but thereafter it increases more rapidly in efficiency as the temperature rises, than does the proton–proton reaction.[31] Above approximately 1.7×107 K, the CNO cycle becomes the dominant source of energy. This temperature is achieved in the cores of main-sequence stars with at least 1.3 times the mass of the Sun.[32] The Sun itself has a core temperature of about 1.57×107 K.[33]: 5  As a main-sequence star ages, the core temperature will rise, resulting in a steadily increasing contribution from its CNO cycle.[25]

Helium fusion

Main sequence stars accumulate helium in their cores as a result of hydrogen fusion, but the core does not become hot enough to initiate helium fusion. Helium fusion first begins when a star leaves the

red giant branch after accumulating sufficient helium in its core to ignite it. In stars around the mass of the Sun, this begins at the tip of the red giant branch with a helium flash from a degenerate helium core, and the star moves to the horizontal branch where it burns helium in its core. More massive stars ignite helium in their core without a flash and execute a blue loop before reaching the asymptotic giant branch. Such a star initially moves away from the AGB toward bluer colours, then loops back again to what is called the Hayashi track. An important consequence of blue loops is that they give rise to classical Cepheid variables, of central importance in determining distances in the Milky Way and to nearby galaxies.[34]: 250  Despite the name, stars on a blue loop from the red giant branch are typically not blue in colour but are rather yellow giants, possibly Cepheid variables. They fuse helium until the core is largely carbon and oxygen. The most massive stars become supergiants when they leave the main sequence and quickly start helium fusion as they become red supergiants. After the helium is exhausted in the core of a star, helium fusion will continue in a shell around the carbon–oxygen core.[20][24]

In all cases, helium is fused to carbon via the triple-alpha process, i.e., three helium nuclei are transformed into carbon via 8Be.[35]: 30  This can then form oxygen, neon, and heavier elements via the alpha process. In this way, the alpha process preferentially produces elements with even numbers of protons by the capture of helium nuclei. Elements with odd numbers of protons are formed by other fusion pathways.[36]: 398 

Reaction rate

The reaction rate density between species A and B, having number densities nA,B, is given by:

where k is the reaction rate constant of each single elementary binary reaction composing the nuclear fusion process:
here, σ(v) is the cross-section at relative velocity v, and averaging is performed over all velocities.

Semi-classically, the cross section is proportional to , where is the de Broglie wavelength. Thus semi-classically the cross section is proportional to .

However, since the reaction involves

quantum tunneling, there is an exponential damping at low energies that depends on Gamow factor EG, giving an Arrhenius equation
:
where S(E) depends on the details of the nuclear interaction, and has the dimension of an energy multiplied for a cross section.

One then integrates over all energies to get the total reaction rate, using the Maxwell–Boltzmann distribution and the relation:

where is the reduced mass.

Since this integration has an exponential damping at high energies of the form and at low energies from the Gamow factor, the integral almost vanished everywhere except around the peak, called Gamow peak,[37]: 185  at E0, where:

Thus:

The exponent can then be approximated around E0 as:

And the reaction rate is approximated as:[38]

Values of S(E0) are typically 10−3 – 103

main-sequence stars give kT of the order of keV.[39]
: ch. 3 

Thus, the limiting reaction in the

abundance of elements in typical stars, the two reaction rates are equal at a temperature value that is within the core temperature ranges of main-sequence stars.[42]

References

Notes

  1. ^ Particle physicist Andrea Pocar points out, "Confirmation of CNO burning in our sun, where it operates at only one percent, reinforces our confidence that we understand how stars work."
  2. ^ "This result therefore paves the way toward a direct measurement of the solar metallicity using CNO neutrinos. Our findings quantify the relative contribution of CNO fusion in the Sun to be of the order of 1 per cent."—M. Agostini, et al.

Citations

  1. ^ a b Hoyle, F. (1946). "The synthesis of the elements from hydrogen". .
  2. ^ a b Hoyle, F. (1954). "On Nuclear Reactions Occurring in Very Hot STARS. I. The Synthesis of Elements from Carbon to Nickel". .
  3. ^ a b Burbidge, E. M.; Burbidge, G. R.; Fowler, W.A.; Hoyle, F. (1957). "Synthesis of the Elements in Stars" (PDF). .
  4. ^ Suess, H. E.; Urey, H. C. (1956). "Abundances of the Elements". .
  5. ^ Clayton, D. D. (1968). Principles of Stellar Evolution and Nucleosynthesis. University of Chicago Press.
  6. ^ Eddington, A. S. (1920). "The internal constitution of the stars".
    PMID 17747682
    .
  7. ^ Eddington, A. S. (1920). "The Internal Constitution of the Stars".
    PMID 17747682
    .
  8. ^ Selle, D. (October 2012). "Why the Stars Shine" (PDF). Guidestar. Houston Astronomical Society. pp. 6–8. Archived (PDF) from the original on 2013-12-03.
  9. ^ Krane, K. S., Modern Physics (Hoboken, NJ: Wiley, 1983), p. 410.
  10. ^ Bethe, H. A. (1939). "Energy Production in Stars".
    PMID 17835673
    .
  11. ..
  12. ^ Clayton, D. D. (1968). Principles of Stellar Evolution and Nucleosynthesis. University of Chicago Press. p. 365.
  13. ^ Clayton, D. D. (2007). "History of Science: Hoyle's Equation".
    S2CID 118423007
    .
  14. ^ Cameron, A. G. W. (1957). Stellar Evolution, Nuclear Astrophysics, and Nucleogenesis (PDF) (Report). Atomic Energy of Canada Limited. Report CRL-41.
  15. ^ Clayton, D. D.; Fowler, W. A.; Hull, T. E.; Zimmerman, B. A. (1961). "Neutron capture chains in heavy element synthesis". .
  16. ^ Seeger, P. A.; Fowler, W. A.; Clayton, D. D. (1965). "Nucleosynthesis of Heavy Elements by Neutron Capture". .
  17. ^ Bodansky, D.; Clayton, D. D.; Fowler, W. A. (1968). "Nucleosynthesis During Silicon Burning". .
  18. ^ Bodansky, D.; Clayton, D. D.; Fowler, W. A. (1968). "Nuclear Quasi-Equilibrium during Silicon Burning". .
  19. ^ Clayton, D. D. (1964). "Cosmoradiogenic Chronologies of Nucleosynthesis". .
  20. ^
  21. Wadsworth Publishing Company
    , 1986), p. 245.
  22. ^
  23. .
  24. ^
  25. ^
  26. .
  27. ^ "Neutrinos yield first experimental evidence of catalyzed fusion dominant in many stars". phys.org. Retrieved 2020-11-26.
  28. ^ Choppin, G. R., Liljenzin, J.-O., Rydberg, J., & Ekberg, C., Radiochemistry and Nuclear Chemistry (Cambridge, MA: Academic Press, 2013), p. 357.
  29. S2CID 227174644
    .
  30. .
  31. ^ Wolf, E. L., Physics and Technology of Sustainable Energy (Oxford, Oxford University Press, 2018), p. 5.
  32. ^ Karttunen, H., Kröger, P., Oja, H., Poutanen, M., & Donner, K. J., eds., Fundamental Astronomy (Berlin/Heidelberg: Springer, 1987), p. 250.
  33. ^ Rehder, D., Chemistry in Space: From Interstellar Matter to the Origin of Life (Weinheim: Wiley-VCH, 2010), p. 30.
  34. ^ Perryman, M., The Exoplanet Handbook (Cambridge: Cambridge University Press, 2011), p. 398.
  35. ^ Iliadis, C., Nuclear Physics of Stars (Weinheim: Wiley-VCH, 2015), p. 185.
  36. ^ "University College London astrophysics course: lecture 7 – Stars" (PDF). Archived from the original (PDF) on January 15, 2017. Retrieved May 8, 2020.
  37. ^ Maoz, D., Astrophysics in a Nutshell (Princeton: Princeton University Press, 2007), ch. 3.
  38. S2CID 16061677
    .
  39. .
  40. ^ Goupil, M., Belkacem, K., Neiner, C., Lignières, F., & Green, J. J., eds., Studying Stellar Rotation and Convection: Theoretical Background and Seismic Diagnostics (Berlin/Heidelberg: Springer, 2013), p. 211.

Further reading

External links