Fact-checked by Grok 2 weeks ago

Stellar nucleosynthesis

Stellar nucleosynthesis is the process by which stars forge chemical elements heavier than and through reactions powered by the immense temperatures and pressures in their interiors. This mechanism accounts for the creation of nearly all elements in the periodic table beyond those primordial light nuclei produced during , fundamentally shaping the chemical composition of the universe. The process unfolds in sequential burning stages as stars evolve, beginning with hydrogen burning in the core where hydrogen nuclei fuse into helium, either via the proton-proton chain in lower-mass stars or the catalytic in more massive ones, releasing energy that sustains the star for billions of years. Once hydrogen is depleted, the core contracts and heats, igniting helium burning at temperatures around 100 million Kelvin, primarily through the that combines three nuclei to form , with subsequent reactions producing and other light elements. In more massive stars (above about 8 solar masses), advanced burning phases follow: carbon burning at roughly 600 million fuses carbon into magnesium, , and sodium; neon burning at 1.2 billion photodisintegrates -20 to produce oxygen and magnesium; oxygen burning at 1.5 billion creates , , and other intermediates; and finally silicon burning at over 3 billion builds up iron-group nuclei like and nickel-56 through a complex network of reactions. These quiescent stages produce elements up to iron, beyond which becomes endothermic and cannot sustain the star, leading to core collapse in massive stars. Explosive nucleosynthesis occurs during cataclysmic events such as core-collapse supernovae or neutron star mergers, where rapid neutron capture—the r-process—and other neutron-rich reactions synthesize heavy elements like gold, platinum, and uranium that are scarce in standard stellar burning. In contrast, the slower s-process (slow neutron capture) operates in asymptotic giant branch stars and red giants, building elements from strontium to lead by gradual neutron addition followed by beta decay. These explosive and asymptotic processes distribute newly formed elements into the interstellar medium via stellar winds and ejecta, enriching subsequent generations of stars and planets. The foundational theory of stellar nucleosynthesis was articulated in the landmark 1957 paper "Synthesis of the Elements in Stars" by E. Margaret Burbidge, G. R. Burbidge, William A. Fowler, and —commonly known as the —which integrated , , and astronomical observations to explain element origins beyond the . Subsequent advancements, including detailed modeling of burning rates and isotopic abundances, have refined these insights, confirming stellar nucleosynthesis as the dominant source of cosmic elements while highlighting ongoing mysteries in heavy-element production sites.

Historical Development

Early Concepts

Ancient Greek philosophers, such as Empedocles and Aristotle, conceptualized the universe as composed of four fundamental elements—earth, air, fire, and water—that could transform into one another through natural processes, providing an early framework for ideas of substance change. Medieval alchemists, building on these notions and influenced by Arabic and Hellenistic traditions, pursued the transmutation of base metals like lead into noble ones such as gold, often through mystical and empirical methods involving distillation and heating, which emphasized the potential mutability of matter. Although these alchemical pursuits focused on terrestrial substances and lacked direct application to celestial bodies, they fostered a cultural and intellectual environment receptive to concepts of elemental transformation that would later inform scientific theories of cosmic origins. In the , amid advances in , British chemist Prout proposed in 1815 that all chemical elements derive from as the primordial building block, hypothesizing that atomic weights were whole-number multiples of hydrogen's mass (approximately 1), implying a process of aggregation from a single protyle or primary matter. This Prout's hypothesis, though challenged by precise measurements showing non-integral atomic weights, represented a speculative shift toward viewing as the universe's foundational element, prefiguring ideas of elemental synthesis from simpler components. Such 19th-century speculations by chemists bridged alchemical legacy with emerging physical sciences, setting the stage for astrophysical applications. Turning to the early 20th century, British astrophysicist , in his 1920 presidential address to the British Association for the Advancement of Science, posited that stellar energy originates from the of four nuclei into one nucleus, releasing energy via mass conversion as per Einstein's E=mc², with roughly 0.7% of the mass lost per reaction. 's proposal, supported by Francis Aston's mass defect measurements and Ernest Rutherford's experiments, addressed the longevity of stars beyond gravitational contraction models but did not detail broader element production, focusing instead on energy generation from . By 1946, Fred Hoyle advanced these ideas in a seminal paper, arguing that helium is synthesized through thermonuclear reactions in the cores of ordinary stars, where high temperatures enable proton captures and subsequent fusions, and that further processes build heavier elements up to iron. Hoyle's work emphasized stellar interiors as active sites for elemental building, predicting specific nuclear pathways without relying on cosmic ray spallation alone. These hypotheses marked the transition toward integrating nuclear physics with stellar evolution.

Key Discoveries and Models

In 1938, and Charles Critchfield proposed the proton-proton chain as the primary mechanism for hydrogen fusion in stars of , calculating the reaction rates for the initial deuteron formation step that initiates the chain, thereby providing a theoretical basis for energy generation in low-mass stars. This model addressed the need for a thermonuclear process capable of sustaining stellar under the temperature and density conditions inferred from theory. Building on this, in 1939 extended the framework by proposing the carbon-nitrogen-oxygen (, a catalytic fusion process where carbon, nitrogen, and oxygen isotopes facilitate hydrogen-to-helium conversion in more massive stars, dominating energy production at higher core temperatures above those required for the proton-proton chain. Bethe's analysis demonstrated that the CNO cycle's efficiency increases rapidly with temperature, explaining the observed mass-luminosity relation for main-sequence stars more massive than . Experimental progress followed in 1952 when Edwin Salpeter refined the proton-proton reaction rate using improved theoretical approximations for the cross-section, enhancing the accuracy of energy production predictions in stellar cores and validating Bethe and Critchfield's earlier estimates. This work marked a pivotal advancement in quantifying the slow initial step of the chain, which limits the overall fusion rate. The landmark 1957 paper by E. Margaret Burbidge, G. R. Burbidge, William A. Fowler, and —known as the B²FH paper—synthesized data with astrophysical observations to outline stellar nucleosynthesis processes from through iron-peak elements, proposing that successive burning stages in massive build heavier nuclei up to iron via equilibrium reactions. Their synthesis incorporated rapid and slow neutron-capture processes for elements beyond iron, linking stellar interiors to cosmic abundance patterns and establishing nucleosynthesis as a core component of stellar theory. During the 1960s, computational advances enabled the integration of detailed nucleosynthesis reaction networks into models, solving equations alongside generation and composition changes, as exemplified by Iben's calculations tracing a 1.25 star from the through burning. These models revealed how convective mixing and mass loss influence element synthesis, providing quantitative predictions for stellar yields that aligned with spectroscopic observations. Observational evidence for these mechanisms emerged from detections starting in the late 1960s, notably the led by Raymond Davis, which confirmed the presence of s produced by proton-proton chain reactions in the Sun's core but measured only about one-third of the predicted flux (∼2.56 units versus ∼7.5 predicted), leading to the solar neutrino problem. This discrepancy challenged stellar models until resolved in the early 2000s by the discovery of flavor oscillations, validating the underlying processes while revealing new physics in behavior.

Fundamental Principles

Nuclear Reactions in Stars

Stellar nucleosynthesis encompasses the of chemical elements heavier than through a series of reactions occurring in the interiors of . These processes transform lighter nuclei into progressively heavier ones, releasing energy that supports and drives evolution. The primary types of nuclear reactions involved include , which builds light elements by combining nuclei such as protons and helium isotopes; , which assembles heavy elements beyond iron by successive addition of neutrons to seed nuclei; and , which breaks down unstable nuclei under intense gamma radiation, often facilitating equilibrium adjustments in reaction networks. reactions dominate the synthesis of elements up to iron, while processes, including slow () and rapid (r-process) variants, account for much of the heavier isotopic abundance. plays a key role in explosive environments, reversing some captures to produce rare isotopes. In stellar plasmas, positively charged nuclei must overcome the Coulomb barrier—the electrostatic repulsion arising from their like charges—to approach close enough for the strong nuclear force to bind them. Quantum mechanical tunneling provides a finite probability for nuclei to penetrate this barrier despite insufficient classical energy, enabling reactions at the lower temperatures and densities found in stellar cores. The energetics of these reactions are characterized by the Q-value, defined as the difference between the total rest mass energy of reactants and products, converted via E = mc^2. Exothermic reactions, with positive Q-values, release this difference, powering the ; for instance, the initial step of hydrogen illustrates this: ^1\mathrm{H} + ^1\mathrm{H} \rightarrow ^2\mathrm{H} + e^+ + \nu_e with a Q-value of 1.442 MeV per reaction. Such energy releases sustain the high temperatures and densities required for ongoing .

Thermonuclear Conditions

Stellar requires extreme thermonuclear conditions within the interiors of stars, where high temperatures overcome the between nuclei, enabling reactions. These conditions vary significantly across and mass, with central temperatures ranging from approximately $10^7 K during hydrogen burning in main-sequence stars to up to $10^9 K in the silicon-burning shells of massive stars. Densities in these regions evolve dramatically, starting at around 100 g/cm³ in the helium-exhausted core of intermediate-mass stars and reaching $10^6 g/cm³ or higher in the dense silicon shells during late evolutionary stages. In stellar cores, pressure support arises primarily from the in the relatively low-density envelopes and outer core regions, where P = \frac{\rho k T}{\mu m_H} balances thermal motion against gravity, with \rho as density, T as , \mu as mean molecular weight, and m_H as mass. However, as core contraction increases density beyond $10^4–$10^5 g/cm³, becomes dominant, providing support independent of via the Fermi-Dirac of degenerate electrons, which prevents further collapse in white dwarfs and influences ignition conditions in evolved . This degeneracy is particularly relevant in the cores of with masses between 1 and 10 masses, where it can suppress carbon ignition until temperatures rise sufficiently. Hydrostatic equilibrium governs the stellar interior, ensuring that the gravitational force per unit volume, -\frac{G m(r) \rho(r)}{r^2}, is balanced by the pressure gradient \frac{dP}{dr}, which defines the scale height and confines reaction zones to central regions where conditions are most extreme. Energy transport mechanisms, including radiative diffusion in stable layers and convective overturning in unstable regions, shape the temperature and density profiles, with radiative transport dominating in the radiative cores of low-mass stars and convection prevailing in the cores of massive stars to efficiently carry fusion-generated heat outward. Convection, driven by superadiabatic gradients, mixes isotopes and fusion products across convective zones, such as the helium-burning core in red giants, thereby influencing the composition available for subsequent burning stages without which nucleosynthesis yields would be limited to isolated shells.

Hydrogen Fusion Processes

Proton-Proton Chain

The proton-proton (pp) chain is the principal pathway for converting to in the cores of main-sequence stars with masses below approximately 1.3 masses, providing the bulk of their energy output through thermonuclear . This dominates in cooler stellar interiors where temperatures range from about 4 to 15 million , as opposed to hotter massive stars that favor catalytic cycles. The chain was theoretically formulated by and Charles L. Critchfield in 1938, identifying the weak interaction-mediated of protons as the initiating step to overcome the electrostatic barrier between positively charged nuclei. The overall net reaction for the pp chain is the of four protons into one , releasing positrons, neutrinos, and : 4\, ^{1}\mathrm{H} \rightarrow \, ^{4}\mathrm{He} + 2\, e^{+} + 2\, \nu_{e} + 26.73\,\mathrm{MeV}. Of this , approximately 26.2 MeV is retained in the as of particles and gamma rays, while 0.53 MeV is carried away by neutrinos on average per helium formed. The chain proceeds through three branches—ppI, ppII, and ppIII—arising from competing reactions involving nuclei, with branching ratios determined by temperature, density, and nuclear cross sections in the stellar core. In , according to standard solar models, the ppI branch accounts for roughly 70% of the production, ppII for about 30%, and ppIII for less than 0.01%. These ratios reflect the relative rates of the key branching points, primarily the competition between channels. All branches share the initial steps, which establish the rate-limiting bottleneck of . The first is the weak of two protons: ^{1}\mathrm{H} + ^{1}\mathrm{H} \rightarrow \, ^{2}\mathrm{H} + e^{+} + \nu_{e}, with a Q-value of 1.442 MeV and maximum energy of 0.423 MeV. This step has an extremely low cross section, on the order of 10^{-18} at energies, due to the need for one proton to undergo during the interaction, making it the primary bottleneck that governs the overall rate in low-mass stars. The subsequent rapidly forms : ^{2}\mathrm{H} + ^{1}\mathrm{H} \rightarrow \, ^{3}\mathrm{He} + \gamma, releasing 5.494 MeV. The branching occurs at the fate of ^{3}\mathrm{He}, with the ppI path involving direct fusion of two such nuclei: ^{3}\mathrm{He} + ^{3}\mathrm{He} \rightarrow \, ^{4}\mathrm{He} + 2\, ^{1}\mathrm{H}, Q = 12.859 MeV, completing the cycle without additional neutrinos beyond the initial pair. This branch dominates at solar temperatures because the ^{3}\mathrm{He} + ^{3}\mathrm{He} rate exceeds that of ^{3}\mathrm{He} capture by ambient ^{4}\mathrm{He}. The ppII branch diverts when ^{3}\mathrm{He} captures ^{4}\mathrm{He}: ^{3}\mathrm{He} + ^{4}\mathrm{He} \rightarrow \, ^{7}\mathrm{Be} + \gamma, Q = 1.587 MeV. The ^{7}\mathrm{Be} then undergoes electron capture: ^{7}\mathrm{Be} + e^{-} \rightarrow \, ^{7}\mathrm{Li} + \nu_{e}, emitting a monoenergetic neutrino at 0.862 MeV (90%) or 0.384 MeV (10%), followed by: ^{7}\mathrm{Li} + ^{1}\mathrm{H} \rightarrow 2\, ^{4}\mathrm{He}. Q = 17.348 MeV. The ppIII branch shares the ^{7}\mathrm{Be} production but branches via proton capture: ^{7}\mathrm{Be} + ^{1}\mathrm{H} \rightarrow \, ^{8}\mathrm{B} + \gamma, Q = 0.137 MeV, a rare event due to its low cross section. The ^{8}\mathrm{B} promptly decays: ^{8}\mathrm{B} \rightarrow \, ^{8}\mathrm{Be} + e^{+} + \nu_{e}, with neutrinos up to 15.0 MeV, and ^{8}\mathrm{Be} subsequently fissions into two ^{4}\mathrm{He} nuclei. This branch is negligible in energy contribution but significant for high-energy neutrino detection. The pp chain produces distinct neutrino species that escape the star unimpeded, providing probes of core conditions. In standard solar models, the predicted fluxes at are dominated by pp neutrinos at $6.0 \times 10^{10} cm^{-2} s^{-1}, reflecting the total burning rate, with ^{7}Be neutrinos at $5.0 \times 10^{9} cm^{-2} s^{-1} (tracing the ppII branch) and ^{8}B at $5.2 \times 10^{6} cm^{-2} s^{-1} (ppIII). These fluxes have been increasingly confirmed by experiments like Borexino, validating the model's inputs, including cross sections extrapolated from and for the weak p-p step. The chain's efficiency and neutrino signatures underscore its role in sustaining and in solar-type stars.

CNO Cycle

The CNO cycle represents a key mechanism of burning in stars, where carbon, , and oxygen isotopes serve as catalysts to facilitate the of four protons into one , releasing primarily through gamma rays and positrons. Proposed by in 1939 as an alternative to the proton-proton chain for hotter stellar interiors, this cycle becomes the dominant source in main-sequence stars with masses exceeding approximately 1.3 solar masses, where core temperatures surpass those required for the pp-chain to compete effectively. The principal branch, known as the CNO-I cycle, operates through a sequence of proton captures, beta decays, and an alpha emission that regenerates the initial carbon seed. The cycle begins with the capture of a proton by ^{12}C, producing ^{13}N, which undergoes positron emission to form ^{13}C; this is followed by another proton capture to yield ^{14}N. Subsequent steps involve proton addition to ^{14}N forming ^{15}O, which beta-decays to ^{15}N, and finally, proton capture on ^{15}N releasing an alpha particle to return to ^{12}C. These reactions, while often summarized in six main steps, encompass twelve individual nuclear processes when accounting for the detailed captures, decays, and emissions closing with the key beta decay ^{15}O \to ^{15}N + e^+ + \nu_e. Minor branches of the CNO cycle, including CNO-II, CNO-III, and CNO-IV, arise from alternative reaction paths at higher temperatures or densities, leading to the production of neon and fluorine isotopes as byproducts. For instance, the CNO-II branch involves leakage from ^{15}N via proton capture to ^{16}O, which can further react to form ^{17}F and ultimately contribute to ^{18}O accumulation; CNO-III and CNO-IV extend this by incorporating ^{18}O and ^{20}Ne seeds, respectively, generating isotopes such as ^{19}Ne, ^{19}F, and ^{21}Ne through proton captures and beta decays that bypass the main cycle's closure. These branches are less efficient than CNO-I but become relevant in massive stars, enhancing the synthesis of light elements beyond helium. The CNO cycle is temperature-sensitive, with its reaction rates increasing rapidly above 15 million Kelvin (T > 1.5 \times 10^7 K), making it negligible in cooler cores like the Sun's but dominant in more massive stars where it supplies over 90% of fusion energy. The net reaction mirrors that of the pp-chain—4^{1}H \to ^{4}He + 2e^+ + 2\nu_e + 26.7 MeV—but relies on trace amounts of CNO seed nuclei (typically 0.1-1% of hydrogen mass fraction) that are recycled without net consumption. In solar models, the cycle contributes only about 1% to the Sun's total energy output due to the lower core temperature, resulting in a predicted low flux of CNO neutrinos; however, the Borexino experiment confirmed their detection in 2020, providing direct evidence of CNO burning in the Sun and validating standard solar composition models, with final results in 2023 measuring the flux at Earth as $6.7^{+1.2}_{-0.8} \times 10^{8} cm^{-2} s^{-1} after neutrino flavor conversion.

Helium and Light Element Fusion

Triple-Alpha Process

The is the primary mechanism by which is fused into carbon in the cores of stars with masses between approximately 0.5 and 8 solar masses during their phase. This occurs under thermonuclear conditions where temperatures reach about $10^8 K and densities are around $10^5 g/cm³, allowing alpha particles to overcome their with sufficient probability. The overall is $3 ^4\mathrm{He} \rightarrow ^{12}\mathrm{C} + \gamma, releasing 7.3 MeV of per , which is carried away primarily as and supports the star's output during helium burning. The process proceeds in two steps: first, two ^4\mathrm{He} nuclei fuse to form an unstable ^8\mathrm{Be} intermediate state with a half-life of about $10^{-16} s, followed by the capture of a third ^4\mathrm{He} to produce ^{12}\mathrm{C}. Without enhancements, the reaction rate would be negligible at stellar temperatures due to the short lifetime of ^8\mathrm{Be}. However, the rate is dramatically increased by the Hoyle resonance, an excited state of ^{12}\mathrm{C} at 7.65 MeV above the ground state, which lies just 300 keV above the energy threshold for the ^8\mathrm{Be} + ^4\mathrm{He} channel and allows resonant capture at relevant thermal energies. This resonance was predicted on astrophysical grounds to explain observed carbon abundances before its experimental confirmation. The reaction rate r for the is given by r = \frac{ \langle \sigma v \rangle n_{\mathrm{He}}^3 }{6}, where n_{\mathrm{He}} is the of nuclei and \langle \sigma v \rangle is the velocity-averaged cross section, with the factor of 1/6 accounting for the indistinguishability of the three identical particles. The \langle \sigma v \rangle is highly sensitive to due to the , peaking near $10^8 and enabling efficient carbon production. This process generates the primary cosmic abundance of carbon, independent of prior , as the carbon yield scales directly with the initial content in the star. Subsequent alpha captures on ^{12}\mathrm{C} can then produce oxygen-16.

Alpha-Chain Reactions

The alpha-chain reactions, also known as the alpha ladder or alpha process, involve sequential captures of helium-4 nuclei (alpha particles) onto carbon-12 seed nuclei produced via the triple-alpha process, building up intermediate-mass elements toward the iron peak during quiescent or shell helium burning in stars. These reactions dominate in the helium-rich environments of massive star cores and asymptotic giant branch (AGB) star shells, where temperatures and densities enable alpha captures over competing photodisintegration. The process contributes significantly to the production of alpha elements like oxygen, neon, magnesium, silicon, and sulfur, which are key to understanding stellar yields and galactic chemical evolution. The chain begins with the reaction ^{12}\mathrm{C} + ^{4}\mathrm{He} \rightarrow ^{16}\mathrm{O} + \gamma, which occurs at temperatures of approximately $10^{8} K in the helium-burning regions of stars with masses above about 0.5 masses. This reaction determines the core carbon-to-oxygen ratio, influencing subsequent evolution and , with its rate remaining a major uncertainty due to experimental challenges at Gamow-peak energies. Successive captures follow at progressively higher temperatures: ^{16}\mathrm{O} + ^{4}\mathrm{He} \rightarrow ^{20}\mathrm{Ne} + \gamma at around $2 \times 10^{8} K, ^{20}\mathrm{Ne} + ^{4}\mathrm{He} \rightarrow ^{24}\mathrm{[Mg](/page/MG)} + \gamma, ^{24}\mathrm{[Mg](/page/MG)} + ^{4}\mathrm{He} \rightarrow ^{28}\mathrm{[Si](/page/Si)} + \gamma, and ^{28}\mathrm{[Si](/page/Si)} + ^{4}\mathrm{He} \rightarrow ^{32}\mathrm{[S](/page/%s)} + \gamma at temperatures rising to $3-4 \times 10^{8} K. These steps proceed efficiently under hydrostatic conditions in helium shells, releasing energy that supports the star against collapse while synthesizing elements essential for later burning stages. Branching in the alpha chain arises from competition between alpha capture and beta decay of unstable intermediates, allowing leakage to the proton-rich side of the isotope chart and producing minor isotopes via pathways like proton captures or alternative decays. These branches are temperature-dependent, with higher rates favoring the main alpha path, and their uncertainties affect predictions of neon and magnesium isotopic distributions in stellar models. In AGB stars of intermediate mass (roughly 2-8 solar masses), alpha-chain reactions occur during recurrent thermal pulses in the helium shell, converting helium into carbon and oxygen that are partially mixed outward via third dredge-up episodes, enhancing surface abundances and yields of alpha elements released into the interstellar medium. This process shapes the chemical evolution of low-metallicity populations, where AGB stars provide a primary source of these elements before supernova contributions dominate. Observational yields from such stars match models incorporating alpha-chain efficiencies, confirming their role in populating the alpha-element plateau in galactic abundances. Mixing between the helium-burning shell, rich in ^{12}\mathrm{C} from alpha chains, and the overlying hydrogen-burning envelope, processed by the to elevate ^{13}\mathrm{C}, lowers the surface ^{12}\mathrm{C}/^{13}\mathrm{C} ratio to values around 10-20 in red giants and AGB stars, serving as a diagnostic of extra-mixing mechanisms like meridional circulation or convective overshoot. This , observed in planetary nebulae and cluster giants, reflects partial dilution of pristine helium-burning products with CN-cycled material, with lower ratios indicating deeper penetration of mixing during the AGB phase.

Advanced Core Burning Stages

Carbon and Neon Burning

Following the exhaustion of helium in the cores of massive with initial masses greater than approximately 8 solar masses, the core contracts under gravitational influence, raising temperatures to around 6 × 10^8 and densities of about 10^5 g cm^{-3}, sufficient to ignite carbon . This stage marks the onset of advanced burning, where nuclei, accumulated from prior helium , begin fusing. The dominant carbon fusion reactions are ^{12}\text{C} + ^{12}\text{C} \to ^{20}\text{Ne} + ^{4}\text{He} (releasing about 4.6 MeV) and ^{12}\text{C} + ^{12}\text{C} \to ^{23}\text{Na} + p (releasing about 2.4 MeV), with the former branch contributing roughly twice as much to energy generation as the latter under typical stellar conditions. These processes primarily produce ^{20}\text{Ne}, along with lesser amounts of ^{23}\text{Na} and, through subsequent alpha captures, isotopes of magnesium such as ^{24}\text{Mg}. Carbon burning occurs convectively in the core due to the high energy release and temperature sensitivity of the reactions, leading to a well-mixed region where neon accumulates as the primary ash. The duration of core carbon burning varies with stellar mass but is generally short on evolutionary timescales, lasting from several hundred years in stars around 20 M_⊙ to about 10^3 years in lower-mass massive stars like 8 M_⊙, driven by the rapid consumption of carbon fuel. In stars of intermediate mass (roughly 4–8 M_⊙), the carbon-oxygen core becomes electron-degenerate prior to ignition, resulting in a thermonuclear runaway known as the carbon flash, where temperatures spike rapidly without significant pressure increase due to degeneracy pressure. Subsequent to carbon exhaustion, the neon-rich core contracts further, reaching temperatures of approximately 1.5 × 10^9 and similar densities, initiating burning. This phase begins with the ^{20}\text{Ne} + \gamma \to ^{16}\text{O} + ^{4}\text{He}, which is endothermic but followed by exothermic alpha captures such as ^{16}\text{O} + ^{4}\text{He} \to ^{20}\text{Ne} + \gamma and further captures leading to ^{24}\text{}. The net effect converts into magnesium isotopes, primarily ^{24}\text{}, with burning proceeding convectively and on an even shorter timescale of roughly 1 year in typical massive stars. This stage builds the neon-magnesium core essential for later evolutionary phases in massive stars.

Oxygen and Silicon Burning

Oxygen burning represents a critical late-stage hydrostatic fusion process in the cores of massive stars, occurring when central temperatures reach approximately 1.5–2 × 10^9 K. At these conditions, oxygen nuclei, primarily ^{16}O accumulated from prior and burning phases, fuse primarily via the reaction ^{16}\text{O} + ^{16}\text{O} \rightarrow ^{28}\text{Si} + ^{4}\text{He}, with additional channels producing (e.g., ^{32}S), (e.g., ^{36}Ar), and lighter elements such as and calcium. This exothermic process releases significant energy, supporting the star against for a limited duration, typically on the order of hours to months depending on the and core density. As oxygen fuel is depleted, the core contracts further, raising temperatures to around 3 × 10^9 K and initiating burning. Unlike earlier fusion stages dominated by a few key reactions, burning involves a dense network of proton-capture ((p,γ)) and alpha-capture ((α,γ)) processes operating in quasi-equilibrium, where forward and reverse reactions balance to favor the most stable nuclei. Starting from the , , and produced in oxygen burning, this network progressively builds heavier isotopes through intermediate steps, ultimately yielding iron-group elements such as ^{56} and ^{56}, which lie near the peak of . The phase unfolds rapidly, often lasting mere seconds to hours in for the most advanced stages, though convective shell burning can extend the overall duration to days. The culmination of silicon burning results in the formation of an inert iron core, where subsequent reactions become endothermic, absorbing rather than releasing . This energy deficit triggers unchecked core contraction and collapse, marking the prelude to a explosion. The silicon- and iron-group nuclei produced serve as seed material for additional heavy-element synthesis during the ensuing explosive .

Heavy Element Production

Slow Neutron Capture Process

The slow neutron capture process, commonly known as the , operates in stellar environments where neutrons are captured by nuclei at a rate slower than the typical beta-decay timescales of the resulting unstable isotopes, allowing the to decay back toward before the next capture occurs. This mechanism was first proposed as a primary pathway for synthesizing heavy elements in stars. The s-process predominantly takes place in low- to intermediate-mass (AGB) stars, during their thermal pulsation phases in the helium-burning shell. There, neutrons are primarily supplied by two reactions: the ^{13}\mathrm{C}(\alpha,n)^{16}\mathrm{O} reaction, which activates during the interpulse periods when accumulated in a thin is exposed to helium burning, and the ^{22}\mathrm{Ne}(\alpha,n)^{25}\mathrm{Mg} reaction, which provides a secondary, more intense but shorter-lived source during convective thermal pulses at higher temperatures. These sources produce neutron densities on the order of $10^7 cm^{-3}, enabling gradual buildup without overwhelming the decay processes. Starting from iron-peak seed nuclei (around A ≈ 56), the proceeds via successive captures and intervening decays, progressively forming heavier nuclei up to lead (Pb) and (Bi) near the end of the stable isotope chain at A ≈ 209. Branching occurs at neutron-rich unstable isotopes where the competition between and determines the path, with captures favored at lower densities to to the valley of beta stability. A key feature of the is the occurrence of bottlenecks at closures, particularly magic neutron numbers like N = 82 (around the Sr-Y-Zr region), where the cross-section drops sharply due to the filled , causing an accumulation of isotopes until or alternative paths proceed. These bottlenecks shape the abundance distribution, with enhanced production at such points. In the classical r/s-process isotope chart, the s-process path traces a curve that hugs the line of stable nuclides, deviating only at branching points and showing distinct s-only isotopes (those not produced by rapid processes) shaded accordingly; this σ(r/s) plot highlights the s-process signature by comparing solar abundances to pure s-process predictions. Overall, the s-process contributes roughly 50% of the solar system's abundances for elements heavier than iron up to , with the remainder from other mechanisms like the neutron capture process, which operates under higher neutron fluxes.

Rapid Neutron Capture Process

The rapid neutron capture process, commonly known as the r-process, is a key nucleosynthesis pathway that accounts for the production of roughly half of the elements heavier than iron in the universe, particularly those with neutron-rich isotopes. This process involves the successive capture of free s by seed nuclei—typically iron-group elements or lighter —in environments characterized by extreme neutron densities exceeding $10^{20} neutrons per cm³. Unlike slower capture mechanisms, the r-process proceeds at rates where neutron captures outpace decays by factors of 100 or more, enabling the formation of highly unstable, neutron-excess nuclei far from . The seminal conceptualization of the r-process was introduced in the 1957 B2FH review, which outlined its role in explosive astrophysical events. During the active phase, a seed nucleus undergoes 10 to 100 rapid neutron captures, building mass numbers up to and beyond A > 209, where the resulting superheavy nuclei become susceptible to fission. This sequence drives the r-process path deep into the neutron-rich region of the nuclear chart, deviating significantly from the valley of beta stability toward the neutron drip line, where neutron separation energies drop below 1 MeV. The path's position, often 10 to 20 neutrons beyond stable isotopes, depends critically on nuclear masses, fission barriers, and beta-decay rates for exotic species, which are probed through facilities like FRIB and GSI. Once the neutron density falls (freeze-out), the highly neutron-rich isotopes decay via beta emission chains, shifting toward stable nuclides and imprinting the observed abundance pattern with peaks at A ≈ 80, 130, and 195. The r-process path's deviation from stability necessitates advanced nuclear theory to predict yields accurately, as uncertainties in far-off-stability properties can alter peak positions by several mass units. The primary site for the r-process is the merger of compact binary systems, particularly neutron star-neutron star collisions, which eject neutron-rich material at velocities of 0.1–0.3c and temperatures around 1–10 GK, ideal for rapid captures. This scenario was observationally validated by the event , detected on August 17, 2017, by and , whose electromagnetic counterpart—a —exhibited spectral features from r-process decay products like and lanthanides, confirming heavy element synthesis in the merger ejecta of about 0.05 masses. Earlier candidates like core-collapse supernovae remain debated due to insufficient production in standard models, but neutron star mergers now dominate theoretical yields for galactic enrichment. The r-process abundance distribution features a prominent third peak near A ≈ 195, populating elements such as (A=197), , and , while extending to actinides including through subsequent captures. For nuclei exceeding A ≈ 209, —either spontaneous, neutron-induced, or beta-delayed—plays a crucial role, fragmenting them into two lighter daughters (typically A ≈ 110 and A ≈ 130) that re-enter the capture flow, a recycling mechanism termed fission cycling. This process moderates overproduction of actinides and bolsters the second and third peaks, with cycling efficiency depending on fission barriers around 6–8 MeV for superheavies. Post-freeze-out, the beta-decay chains from these neutron-rich progenitors release the final r-process elements, with half-lives ranging from milliseconds to years, powering emission and contributing to products. The r-process thus complements slower pathways by synthesizing the most , neutron-rich heavy nuclides.

Explosive Nucleosynthesis

Core-Collapse Supernovae

Core-collapse supernovae (CCSNe) represent a critical site for explosive nucleosynthesis, occurring in the final stages of massive stars with initial masses exceeding 8 solar masses, where the collapse of an iron core triggers a rebounding that propagates outward through the stellar . This shock induces rapid heating and compression in the overlying layers, enabling nuclear reactions at temperatures and densities unattainable during hydrostatic evolution, thereby synthesizing a broad range of elements from oxygen-burning products up to and beyond the iron peak. Unlike the quasi-static burning phases, the explosive environment is characterized by short timescales (seconds to minutes) and high entropies, leading to incomplete and distinct isotopic signatures. A key aspect of explosive burning in CCSNe involves the shock-heated oxygen and -rich shells, where temperatures reach 3–5 GK, facilitating explosive oxygen burning that primarily produces silicon-28, sulfur-32, argon-36, and calcium-40 through alpha-capture reactions on lighter seeds. In the shell, explosive silicon burning at similar temperatures drives quasi-equilibrium reactions toward the iron-peak elements (Sc to Zn), but incomplete silicon burning in marginally heated regions (T < 4 GK) results in an overproduction of intermediate-mass nuclei like silicon-28 and sulfur-32 relative to full equilibrium products, contributing to the observed abundance patterns in supernova remnants. These processes account for much of the synthesis of elements between neon and nickel, with the Fe-peak dominated by nickel-56 decaying to cobalt-56 and iron-56. The ν-process, a neutrino-induced spallation mechanism, operates concurrently in the helium shell and outer layers, where high-flux neutrinos from the proto-neutron star interact with CNO seed nuclei to produce light, neutron-poor isotopes such as , , and via reactions like ^{12}C(ν, p)^{11}B and ^{16}O(ν, α)^{13}C followed by further captures. First detailed by Woosley et al. (1990), this process relies on the intense neutrino luminosity (~10^{51} erg) during the first seconds post-bounce and explains observed abundances of these elements in metal-poor stars without requiring galactic cosmic rays as the sole source. Yields from the ν-process are sensitive to neutrino spectra and progenitor metallicity, typically contributing 10–30% of solar . In neutron-rich ejecta (electron fraction Ye < 0.45), the can occur, particularly in regions with high neutron-to-seed ratios, synthesizing heavy elements beyond the Fe-peak up to thorium and uranium through successive neutron captures on iron-group seeds followed by beta-decays. While binary neutron star mergers are now favored as primary r-process sites, some CCSN models with strong magnetic fields or rapid rotation produce neutron-rich conditions in the inner ejecta, potentially contributing to lighter r-process peaks (A ~ 80–130). Isotopic observations, such as europium enhancements in metal-poor stars, support a mixed contribution from CCSNe in the early Galaxy. Nucleosynthetic yields from CCSNe significantly shape galactic chemical evolution, producing approximately 50% of the solar system's iron inventory through decay of in the ejecta (mean yield ~0.07 M_⊙ per event), with the remainder from Type Ia supernovae. These events also yield substantial titanium and chromium, with isotopic anomalies (e.g., deficits in ^{48}Ti and ^{54}Cr) observed in presolar grains and meteorites, attributed to incomplete burning and neutron capture in the exploding layers. For a typical 15–25 M_⊙ progenitor, explosive yields include ~0.5–1 M_⊙ of oxygen, ~0.1 M_⊙ of silicon, and ~0.05 M_⊙ of iron-peak elements, varying with explosion energy (1–2 × 10^{51} erg). Three-dimensional simulations of CCSNe emphasize the role of neutrino-driven winds emanating from the proto-neutron star, which carry ~0.01–0.1 M_⊙ of material at velocities of 10,000–30,000 km/s and moderate entropies (20–50 k_B per baryon), enabling weak r-process or νp-process nucleosynthesis in proton-rich conditions. These winds, powered by neutrino absorption on free nucleons (e.g., \bar{ν}_e + p → n + e^+), deposit ~10^{51} erg and alter Ye from proton-rich (Ye > 0.5) to potentially neutron-rich values over ~10 seconds, influencing yields of elements like and . Recent multi-dimensional models, incorporating realistic transport, confirm that wind contributions to heavy elements are modest but crucial for resolving discrepancies in light r-process abundances.

Type Ia Supernovae and Novae

Type Ia supernovae arise from the thermonuclear explosion of a carbon-oxygen in a that accretes mass until reaching the , triggering explosive carbon burning throughout the star. This process ignites at densities around 10^9 g cm^{-3}, where carbon fusion rapidly propagates as a or , synthesizing primarily iron-peak elements through nuclear statistical equilibrium at temperatures exceeding 10^9 K. The dominant product is ^{56}, which decays via to ^{56}Co (half-life 77 days) and then to stable ^{56} (half-life 111 days), powering the supernova's through gamma-ray emission and annihilation. These events produce a nearly uniform yield of approximately 0.6 M_\odot of ^{56} per explosion, corresponding to the eventual iron output, which contributes significantly to the iron abundance in the . Classical novae, in contrast, involve explosive hydrogen burning on the surface of an accreting , driven by the accumulation of a hydrogen-rich from a low-mass companion. Ignition occurs at base temperatures around 10^8 , leading to a thermonuclear runaway via the hot , where proton captures on CNO seed nuclei at high temperatures (T > 10^8 ) break out of the standard CNO equilibrium to produce isotopes such as ^{13}C, ^{15}N, and ^{7}Li (via the ^{3}He(\alpha,\gamma)^{7}Be \to ^{7}Li ). These outbursts eject envelopes with masses typically on the order of 10^{-5} M_\odot at velocities of 100-3000 km s^{-1}, enriching the with these light isotopes; models suggest novae contribute up to about 35% of Galactic ^{13}C and ^{15}N abundances. Additionally, oxygen-neon white dwarfs in novae overproduce ^{7}Be, which decays to ^{7}Li, contributing to the primordial lithium problem in some scenarios. Certain models of Type Ia supernovae, particularly delayed-detonation scenarios, predict overproduction of proton-rich p-nuclei (e.g., ^{92}Mo, ^{94}Mo, ^{96}Ru) through gamma-process reactions on seed nuclei in the oxygen-burning zones, potentially explaining 20-50% of solar p-nuclei abundances depending on the explosion strength. While novae primarily affect light elements, Type Ia events play a key role in galactic chemical evolution by uniformly distributing iron-peak material over cosmic time.

Modeling and Observational Evidence

Reaction Rates and Cross-Sections

In stellar nucleosynthesis, the rate at which nuclear reactions proceed is determined by the reaction cross-section \sigma(E), which quantifies the probability of interaction between particles at center-of-mass energy E, averaged over the thermal velocity distribution in the stellar plasma. The cross-section for charged-particle reactions is typically expressed through the astrophysical S-factor, S(E) = E \exp(2\pi\eta) \sigma(E), where \eta is the Sommerfeld parameter accounting for the Coulomb barrier, to isolate the nuclear interaction from electrostatic effects. This formulation facilitates extrapolation to the low energies relevant for stellar interiors, where direct measurements are challenging due to the exponentially suppressed cross-sections. The thermonuclear is obtained by averaging the product of the cross-section and over the Maxwell-Boltzmann velocity distribution of the particles: \langle \sigma v \rangle = \left( \frac{8}{\pi \mu} \right)^{1/2} \frac{1}{(kT)^{3/2}} \int_0^\infty \sigma(E) E \exp\left( -\frac{E}{kT} \right) dE, where \mu is the , k is Boltzmann's constant, and T is the ; this captures the thermal averaging essential for computing energy generation and abundance changes in models. For non-resonant reactions dominated by the tail of the Maxwell-Boltzmann distribution penetrating the , the integrand peaks at the Gamow energy E_0, known as the Gamow peak: E_0 = \left( \frac{\pi \alpha Z_1 Z_2 \hbar c}{\sqrt{2 \mu}} \right)^{2/3} (kT)^{1/3}, which sets the effective energy window for most reactions in hydrostatic stellar burning stages. The width of this peak, \Delta E_0 \approx 4 (kT)^2 / E_g where E_g = 2\pi \alpha Z_1 Z_2 \hbar c / \sqrt{2 \mu} is the Gamow energy scale, determines the sensitivity of the rate to the S-factor near E_0. Resonances, arising from compound nuclear states, can significantly enhance reaction rates if their energies lie within or near the Gamow peak. The contribution from a narrow resonance at energy E_r is characterized by its strength \omega \gamma, defined as \omega \gamma = \frac{2J+1}{(2j_1+1)(2j_2+1)} \Gamma, where J is the total angular momentum of the resonance, j_1 and j_2 are the spins of the incoming particles, and \Gamma is the total width of the resonance; this parameter governs the peak cross-section \sigma_r = \lambda^2 \omega \gamma / (2 \Gamma) and its thermal averaging. In cases where multiple resonances contribute, their strengths are determined through indirect methods like transfer reactions when direct measurements are infeasible at stellar energies. Experimental determinations of cross-sections at or near the Gamow peak are crucial for reducing uncertainties in reaction rates, particularly for light-ion reactions in early burning stages. Underground laboratories, such as the Laboratory for Underground Nuclear Astrophysics () at Gran Sasso, mitigate cosmic-ray backgrounds to measure cross-sections at ultra-low energies relevant to the proton-proton (pp) chain. For instance, has provided direct measurements of the ^3He(\alpha, \gamma)^7Be reaction at solar Gamow peak energies (around 20-30 keV), yielding S-factors with uncertainties below 5%, which refine the predicted solar neutrino flux from the pp chain. Similarly, 's work on ^3He(^3He, 2p)^4He has constrained the ^7Be neutrino branch, impacting models of solar hydrogen burning. For heavier-ion reactions, such as those in advanced burning stages (e.g., ^{12}C + ^{12}C), extrapolations from higher-energy data introduce larger uncertainties, often spanning factors of 2-10 in the reaction rates at stellar temperatures (T_9 ~ 0.5-2, where T_9 = T/10^9 K). These arise from ambiguities in sub-barrier hindrance, resonant contributions, and incomplete knowledge of partial widths, as seen in the ongoing debate over the ^{12}C + ^{12}C rate, where experimental discrepancies lead to variations in predicted carbon-burning lifetimes and subsequent . Theoretical models, including coupled-channels calculations, help quantify these uncertainties but highlight the need for further low-energy data. Such rate uncertainties propagate into simulations, affecting the timing and efficiency of core burning phases.

Stellar Yields and Isotopic Abundances

Stellar yields quantify the net production of chemical elements and isotopes by stars, defined as the mass of a given isotope ejected into the interstellar medium minus the amount initially present in the star, normalized per unit mass of the progenitor star. These yields are computed through detailed stellar evolution models coupled with nuclear reaction networks that track the transformation of over 3000 isotopes via thousands of reactions, spanning from hydrogen burning to explosive endpoints. Such networks incorporate reaction rates from laboratory measurements and theoretical extrapolations, enabling predictions of isotopic outputs from diverse stellar sites including massive stars, asymptotic giant branch (AGB) stars, and supernovae. Galactic chemical evolution (GCE) models integrate these stellar across a range of initial masses and to simulate the temporal and spatial buildup of elemental abundances in galaxies like the . By convolving with the , history, and inflow/outflow processes, GCE frameworks reproduce observed gradients and abundance ratios, such as the gradual enrichment in alpha elements from core-collapse supernovae. Variations in yield tables—arising from differences in treatment or mass-loss prescriptions—can significantly alter predicted evolution, highlighting the need for consistent yield sets in multi-zone simulations. Observational validation of these models relies on comparisons to isotopic ratios in the solar system and primitive meteorites, providing direct tests of yield predictions. For instance, the solar system's ^{16}O/^{18}O ratio of approximately 500 aligns with GCE models that balance oxygen production from massive stars and AGB contributions, though slight discrepancies persist due to uncertainties in low-mass star yields. Presolar grains extracted from meteorites, such as and oxide condensates, exhibit isotopic anomalies—e.g., enrichments in ^{13}C or ^{26}—that match signatures from specific stellar nucleosynthesis episodes, confirming yields from AGB stars and supernovae while revealing grain-specific mixing zones. Despite advances, key incompletenesses limit the accuracy of yield predictions, particularly for proton-rich isotopes produced via the p-process in explosive environments like core-collapse supernovae and Type Ia events. Uncertainties in photon-induced reactions and incomplete network coverage lead to variations in p-nuclide abundances by factors of up to 50%, complicating their integration into GCE models. In AGB stars, the efficiency of the third dredge-up—which mixes carbon and elements from the convective thermal pulse region to the surface—remains poorly constrained due to ambiguities in convective boundary mixing and overshoot parameters, resulting in yield uncertainties exceeding an order of magnitude for elements like carbon and . These gaps underscore the ongoing need for refined stellar interior physics and expanded nuclear data to refine isotopic abundance forecasts.

References

  1. [1]
    DOE Explains...Nucleosynthesis - Department of Energy
    Stars can create nuclei through two processes: either by combining two smaller nuclei (called fusion) or breaking a larger nucleus into multiple nuclei (called ...
  2. [2]
    Basics - Composition - Nucleosynthesis - NASA's Cosmicopia
    May 10, 2012 · Nucleosynthesis is the process where lighter elements fuse into heavier elements in stars, requiring high temperatures. Supernovae also form ...
  3. [3]
    Origin of the Chemical Elements - T. Rauscher & A. Patkos
    An overview of stellar nucleosynthesis including hydrogen, helium, neon, silicon, and explosive burning as well as the basics of the s- and r-process are given ...
  4. [4]
    Philosophy of Chemistry
    Mar 14, 2011 · Aristotle also contributed the first important analyses of this process, distinguishing between transmutation, where one substance overwhelms ...
  5. [5]
    Chemistry - Alchemy, Transmutation, Philosopher's Stone | Britannica
    One of the leading ideas of medieval Arabic alchemy was the theory that all metals were formed of sulfur and mercury in various proportions and that altering ...
  6. [6]
    The Secrets of Alchemy | Science History Institute
    Jan 30, 2013 · The book surveys the history of alchemy from its origins in late antiquity to the present day. It focuses on a few representative characters and ideas.
  7. [7]
    Is all matter made of just one element? | Feature - Chemistry World
    Apr 1, 2015 · Prout's belief that all the elements are made of hydrogen was thus vindicated, though in a way he could never have dreamed of. The non-integral ...Missing: primordial | Show results with:primordial
  8. [8]
    Prout's Hypothesis. Alembic Club Reprint No. 20. Papers by William ...
    $$0.65. Prout's Hypothesis—that the atomic weightsof the elements are multiples of that of hydrogen—dates from the papers which. Prout published anonymously ...
  9. [9]
    [PDF] Eddington and the Problem of Stellar Energy - arXiv
    Eddington was convinced that nuclear processes were the key to unravel the mystery of stellar energy but undecided of whether fusion or annihilation was the.
  10. [10]
    The synthesis of the elements from hydrogen
    (ii) Helium is synthesized by thermonuclear reactions taking place in "normal" stars. (iii) A further process occurs that synthesizes higher elements from ...
  11. [11]
    The Formation of Deuterons by Proton Combination | Phys. Rev.
    The Formation of Deuterons by Proton Combination. H. A. Bethe · C. L. Critchfield.
  12. [12]
    Energy Production in Stars | Phys. Rev.
    It is shown that the most important source of energy in ordinary stars is the reactions of carbon and nitrogen with protons.
  13. [13]
    Nuclear Reactions in the Stars. I. Proton-Proton Chain - ADS
    The rates are calculated for the main reactions making up the "proton-proton chain," whose net effect is the conversion of four hydrogen atoms into one ...
  14. [14]
    Synthesis of the Elements in Stars | Rev. Mod. Phys.
    Synthesis of the Elements in Stars. E. Margaret Burbidge, GR Burbidge, William A. Fowler, and F. Hoyle.
  15. [15]
    [PDF] Synthesis of the Elements in Stars - MIT
    both of the fundamental building blocks.The following question can be asked: Whathas been the history of the matter, on whichwe can make observations, which.Missing: B2FH | Show results with:B2FH
  16. [16]
    [PDF] Stellar Nucleosynthesis - Princeton University
    Insofar as the fuel for these burning stages consists of two- body collisions between nuclei of the same isotope, one might naively expect that atomic.
  17. [17]
    [PDF] thielemann.r-process.sites.pdf - Department of Astrophysical Sciences
    This article addresses three of the four nucleosynthesis processes involved in producing heavy nuclei beyond Fe (with a main focus on the r-process).
  18. [18]
    [PDF] The physics of fusion in stars
    But quantum-mechanical tunneling allows the protons to go. “under” the Coulomb barrier with a probability that is much larger than this, though still.
  19. [19]
    [PDF] Stellar Nucleosynthesis
    Since any reaction branch that completes this must turn 2 protons in 2 neutrons, two neutrinos are also emitted, which carry energy away from the reaction site.
  20. [20]
    [PDF] Nuclear Masses and Mass Excess: Q values for Nuclear Reactions
    The Q value of a nuclear reaction is the difference between the sum of initial reactant masses and final product masses, in energy units.
  21. [21]
    Nuclear Fusion in Protostars | ASTRO 801
    In this multi-step process, six protons fuse together and the product is a helium nucleus and two protons. Here are the steps in equation form: 1H + 1H = 2H + ...
  22. [22]
  23. [23]
  24. [24]
  25. [25]
    (PDF) HYDROSTATIC EQUILIBRIUM OF A STAR - ResearchGate
    Apr 14, 2025 · Hydrostatic equilibrium is a key concept in stellar astrophysics, describing the balance between a star's inward gravitational pull and outward pressure forces.
  26. [26]
    [PDF] Chapter 5 - Energy transport in stellar interiors - Astrophysics
    At high densities and low temperatures, the conductive opacity becomes very small because of the large electron mean free path in a highly degenerate gas. This ...
  27. [27]
  28. [28]
    Solar fusion cross sections. II. The chain and CNO cycles
    Apr 12, 2011 · Instead, this cross section must be calculated from standard weak interaction theory. As in Solar Fusion I, the adopted value and range for ...
  29. [29]
    [PDF] Introduction to Nuclear Astrophysics1 - arXiv
    Nov 20, 2009 · Abstract. In the first lecture of this volume, we will present the basic fundamental ideas regarding nuclear processes occurring in stars.
  30. [30]
    Experimental evidence of neutrinos produced in the CNO fusion ...
    Nov 25, 2020 · Solar neutrinos are detected by Borexino via their elastic scattering off electrons. The total number of detected photons and their arrival ...
  31. [31]
    Final results of Borexino on CNO solar neutrinos | Phys. Rev. D
    Nov 14, 2023 · Borexino delivered the first direct experimental proof of the existence of the CNO cycle in the Sun with a significance of ∼ 7 σ , also ...Article Text · INTRODUCTION · THE BOREXINO EXPERIMENT · CONCLUSIONS
  32. [32]
    Nuclear fusion in dense matter: Reaction rate and carbon burning
    Carbon burning in evolved massive stars takes place at typical densities of ρ ~ 10 5 g cm - 3 and temperatures T ~ ( 6 – 8 ) × 10 8 K. Carbon burning is also ...
  33. [33]
    12C+12C Reaction Rates and the Evolution of a Massive Star - arXiv
    Jun 19, 2024 · We examine the evolution of a 20 M_sun star with the updated 12C+12C reaction rates performing simulations with the MESA (Modules for Experiments for Stellar ...
  34. [34]
    Carbon Burning - an overview | ScienceDirect Topics
    Carbon burning is defined as a nuclear fusion process that occurs in the center of a star after a contraction phase, primarily producing neon-20 from ...
  35. [35]
    A new 12C + 12C nuclear reaction rate: Impact on stellar evolution
    This work presents new 12 C + 12 C reaction rates in the form of numerical tables with associated uncertainty estimations, as well as analytical formulae.
  36. [36]
    3D hydrodynamic simulations of carbon burning in massive stars
    The evolution time-scale is of the order of days (⁠ D a ∼ 1 a n d P e ≫ 1 ⁠). Early simulations of silicon burning (Bazán & Arnett 1997) used a nuclear reaction ...
  37. [37]
    Explosive Ignition of Carbon in Stars of Intermediate Mass - Nature
    Hoyle and Fowler1 have suggested that type I supernovae are the result of an explosion following the ignition of degenerate nuclear fuel in an evolved star.
  38. [38]
    Nucleosynthesis and Chemical Evolution of Oxygen
    Jan 1, 2008 · The first nuclear fuel available to stars is hydrogen, which stars burn into 4He. Stars with the mass of the Sun or less burn hydrogen at ...
  39. [39]
    (PDF) Relation between the 16O(α,γ)20Ne reaction and its reverse ...
    Aug 6, 2025 · The reactions 12 C(α,γ) 16 O [13,14] and 16 O(α,γ) 20 Ne [15, 16] process 12 C and 16 O rates and determine the final 12 C/ 16 O abundance ratio ...
  40. [40]
    Stellar Death - University of Oregon
    Neon burning for 1 year, oxygen burning about 6 months (i.e. very fast on astronomical timescales). At 3 billion degrees, the core can fuse silicon nuclei ...<|separator|>
  41. [41]
    Massive Star Modeling and Nucleosynthesis - Frontiers
    This review intends to summarize the safe grounds and challenges of massive star evolution modeling and nucleosynthesis.
  42. [42]
    High-Mass Stars - Astronomy 1101 - The Ohio State University
    Inert C-O core collapses & heats up; H & He burning into shells; Becomes a Red Supergiant again. End of Blue Supergiant Phase to Second Red Supergiant Phase. ( ...Missing: ranges | Show results with:ranges
  43. [43]
    Populating the periodic table: Nucleosynthesis of the elements
    Feb 1, 2019 · The oxygen left over from helium burning is inert at the temperatures of carbon and neon burning, but at 1 billion K, it too ignites (11, 15).
  44. [44]
    Silicon Burning. II. Quasi-Equilibrium and Explosive Burning
    In all three cases, we find that the important abundances obey quasi-equilibrium for temperatures greater than approximately 3×109 K, with relatively little ...
  45. [45]
    [PDF] Nucleosynthesis During Silicon Burning - Clemson OPEN
    burning, which determines the time scales and energy generation during silicon burning, and which shows how silicon burning accounts for many crucial ...<|control11|><|separator|>
  46. [46]
    [PDF] Lecture 12 Advanced Stages of Stellar Evolution – II Silicon Burning ...
    ~ 7 hours. Shell burning and convection can lengthen this to days to weeks. Page 17. Page 18. Nucleosynthesis. Basically, silicon burning in the star's core ...Missing: seconds | Show results with:seconds
  47. [47]
    Origin of the heaviest elements: The rapid neutron-capture process
    Feb 1, 2021 · The two main processes involving the capture of free neutrons are the slow ( s ) process and the rapid ( r ) process (due to low or high ...
  48. [48]
    [PDF] THE R-PROCESS AND NUCLEOCHRONOLOGY John J. COWAN ...
    This process of heavy-element synthesis involves the progressive buildup of heavier isotopes via neutron captures proceeding on neutron-rich isotopes, ...
  49. [49]
    Spectroscopic identification of r-process nucleosynthesis in ... - Nature
    Oct 16, 2017 · Observations of the transient associated with the gravitational-wave event GW170817 and γ-ray burst GRB 170817A reveal a bright kilonova ...Methods · Fors2 Spectroscopic... · Host Galaxy Emission...Missing: paper | Show results with:paper
  50. [50]
    GW170817$-$the first observed neutron star merger and its kilonova
    Jan 25, 2019 · GW170817 was the first observed neutron star merger, with its emission powered by the r-process, and it marked the beginning of multi-messenger ...Missing: confirmation | Show results with:confirmation
  51. [51]
    Influence of Spontaneous Fission Rates on the r-process ...
    Jun 28, 2022 · We show that the relative contributions of various fission channels in r-process nucleosynthesis depend on the astrophysical conditions and fission models used.
  52. [52]
    [1206.2503] Explosion Mechanisms of Core-Collapse Supernovae
    Jun 12, 2012 · The neutrino-heating mechanism, aided by nonradial flows, drives explosions, albeit low-energy ones, of ONeMg-core and some Fe-core progenitors.
  53. [53]
    [PDF] Lecture 15 Explosive Nucleosynthesis and the r-Process
    As the shock wave passes through the star, matter is briefly heated to temperatures far above what it would have experienced had it burned.
  54. [54]
    CONSTRAINTS ON EXPLOSIVE SILICON BURNING IN CORE ...
    Jul 6, 2015 · ABSTRACT. Measurements of explosive nucleosynthesis yields in core-collapse supernovae provide tests for explosion models. We investigate ...
  55. [55]
    [0707.2187] Nucleosynthesis in Core-Collapse Supernovae and GRB
    Jul 15, 2007 · We review the nucleosynthesis yields of core-collapse supernovae (SNe) for various stellar masses, explosion energies, and metallicities.
  56. [56]
    Neutrino-driven wind simulations and nucleosynthesis of heavy ...
    Jul 11, 2012 · In this review, we summarize our present understanding of neutrino-driven winds from the dynamical and nucleosynthesis perspectives.
  57. [57]
    Comparison between Core-collapse Supernova Nucleosynthesis ...
    Mar 17, 2022 · ... Cr anomalies, while material from the C ashes does not present significant Mg anomalies together with Cr isotopic variations. In all cases ...
  58. [58]
    Nucleosynthesis-relevant conditions in neutrino-driven supernova ...
    After the initiation of the explosion of core-collapse supernovae, neutrinos emitted from the nascent neutron star drive a supersonic baryonic outflow. This ...
  59. [59]
    [2403.12942] The physics of Core-Collapse Supernovae: explosion ...
    Mar 19, 2024 · In this review, we will briefly summarize the state-of-the-art of both 1D and 3D simulations and how they can be employed to study the evolution ...
  60. [60]
    [PDF] arXiv:2111.14795v1 [astro-ph.HE] 29 Nov 2021
    Nov 29, 2021 · Type Ia supernovae (SNe Ia) are well-understood as the thermonuclear explosions of carbon-oxygen white dwarfs (CO WDs)1–3. They produce the ...
  61. [61]
    [PDF] Type Ia supernovae - arXiv
    Jun 5, 2025 · At sufficiently high densities, carbon fusion proceeds all the way to the iron peak, and nuclear statistical equilibrium (NSE) is established ...
  62. [62]
    Nucleosynthesis of Iron-peak Elements in Type-Ia Supernovae
    Healthy 56Ni production about 0.6 solar mass; 2. Consistency of the 55Mn to the galactic chemical evolution; 3. Agreement of 58Ni with the solar abundance. 1.
  63. [63]
    Type Ia supernovae from non-accreting progenitors
    Typical SN Ia luminosities (∼1043 erg s−1) and ejecta velocities (∼104 km s−1) require 56Ni masses and kinetic energies of ∼0.6 M⊙ and ∼1051 erg. ... Branch, D., ...
  64. [64]
    NUCLEOSYNTHESIS IN CLASSICAL NOVAE - IOP Science
    Our estimates show that nova outbursts may account for. 100% of the Galactic abundances of 13C, 15N, and 17O, assuming solar system levels for the mean ...
  65. [65]
    [PDF] arXiv:2006.13589v1 [nucl-ex] 24 Jun 2020
    Jun 24, 2020 · Nova grains show high excesses of isotopes 13C and 15N compared to the solar composition, that can be explained by the hot CNO cycle during ...
  66. [66]
    Hydrodynamic Simulations of Oxygen–Neon Classical Novae as ...
    Feb 20, 2024 · The Amount of 7Li in the Galaxy Produced by Classical. Novae. Our results confirm that ONe novae are also overproducing. 7Be, which decays to ...Missing: yields | Show results with:yields
  67. [67]
    TYPE Ia SUPERNOVAE AS SITES OF THE p-PROCESS - IOP Science
    Beyond Fe, there is a class of 35 proton-rich nuclides, between 74Se and 196Hg, called p-nuclei. They are bypassed by the s and r neutron capture processes ...
  68. [68]
    Gamow peak in thermonuclear reactions at high temperatures
    Apr 3, 2007 · Nuclear reactions in a stellar plasma will occur near energies where the product of velocity distribution and cross section is at maximum. The ...
  69. [69]
    effect of 12C +12C rate uncertainties on the evolution and ...
    Over the last 40 years, the 12C +12C fusion reaction has been the subject of considerable experimental efforts to constrain uncertainties at temperatures ...
  70. [70]
    Theoretical uncertainty quantification for heavy-ion fusion
    Nov 18, 2022 · The purpose of this work is to investigate the uncertainties, originating from the EDF, in the prediction of fusion cross sections from deep sub ...
  71. [71]
    6. Nucleosynthetic Yields — VICE 1.3.1 documentation
    This is the amount of an element that is produced and ejected to the interstellar medium minus that which was already present, in units of the star or stellar ...
  72. [72]
    Type Ia Supernova Nucleosynthesis: Metallicity-dependent Yields
    Aug 22, 2023 · In this work, we calculate complete abundance yields for 39 models of SN Ia explosions, based on three progenitors—a 1.4 M⊙ deflagration ...
  73. [73]
    The Dawes Review 2: Nucleosynthesis and stellar yields of low and ...
    May 1, 2014 · Here we review the stellar evolution and nucleosynthesis for single stars up to ~10Msun from the main sequence through to the tip of the asymptotic giant ...
  74. [74]
    Galactic chemical evolution: stellar yields and the initial mass function
    Other well-known stellar yields are those that include a treatment of stellar rotation, the internal mixing and structural changes resulting from which can lead ...
  75. [75]
    Galactic Chemical Evolution: Stellar Yields and the Initial Mass ...
    May 13, 2015 · We present a set of 144 galactic chemical evolution models applied to a Milky Way analogue, computed using four sets of low and intermediate star ...
  76. [76]
  77. [77]
    [PDF] Supplementary Materials for - NASA Technical Reports Server (NTRS)
    Measurements of 16O/18O throughout the solar system. Isotope ratio. Molecule Reference. Sun. 499±119. O. (53). Jupiter. 243±203. H2O. (57). Venus. 500±26. H2O.
  78. [78]
    Uncertainties in the production of p nuclei in massive stars obtained ...
    Nuclear data uncertainties in the production of p nuclei in massive stars have been quantified in a Monte Carlo procedure.