Topical Review The following article is Open access

Progress on nuclear reaction rates affecting the stellar production of 26Al

, , , , , , , , , , , , , , , , , , , , , , , , , and

Published 28 February 2023 © 2023 The Author(s). Published by IOP Publishing Ltd
, , Citation A M Laird et al 2023 J. Phys. G: Nucl. Part. Phys. 50 033002 DOI 10.1088/1361-6471/ac9cf8

0954-3899/50/3/033002

Abstract

The radioisotope 26Al is a key observable for nucleosynthesis in the Galaxy and the environment of the early Solar System. To properly interpret the large variety of astronomical and meteoritic data, it is crucial to understand both the nuclear reactions involved in the production of 26Al in the relevant stellar sites and the physics of such sites. These range from the winds of low- and intermediate-mass asymptotic giant branch stars; to massive and very massive stars, both their Wolf–Rayet winds and their final core-collapse supernovae (CCSN); and the ejecta from novae, the explosions that occur on the surface of a white dwarf accreting material from a stellar companion. Several reactions affect the production of 26Al in these astrophysical objects, including (but not limited to) 25Mg(p, γ)26Al, 26Al(p, γ)27Si, and 26Al(n, p/α). Extensive experimental effort has been spent during recent years to improve our understanding of such key reactions. Here we present a summary of the astrophysical motivation for the study of 26Al, a review of its production in the different stellar sites, and a timely evaluation of the currently available nuclear data. We also provide recommendations for the nuclear input into stellar models and suggest relevant, future experimental work.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The radioactive nucleus 26Al 29 has gained significant attention in the past decades within several fields of astronomy and planetary science. Its half-life is 0.717(24) Myr [1], which corresponds to a mean lifetime of 1.035 Myr. This value allows the 26Al produced in stars and supernovae, to live, on the one hand, long enough to allow us to trace its abundance back to its creation events, on the other hand, short enough to provide us snapshots of the Galaxy at very specific times, such as today and when the Solar System formed, 4.6 Gyr ago [2]. The fact that its decay produces high-energy γ photons at 1.8 MeV ensures that this radioactive process is observationally accessible via the γ-ray spectrometry performed by satellite observatories (section 1.1). Furthermore, the high-energy γ photons emitted by the decay of 26Al produce a significant amount of heat within rocks that incorporate 26Al at their formation. This is relevant for the first planetesimals that formed in the early Solar System, and potentially for extrasolar planetary systems (section 1.3). Interestingly, we have the samples and the laboratory tools to be able to detect and measure the abundance of 26Al both 'alive' today in the Galactic interstellar medium, as it was potentially deposited inside Earth samples by nearby stellar sources within the past few Myr (section 1.2 [3]), as well as 'extinct', i.e. already fully decayed into 26Mg. In fact, from the measured 26Mg/24Mg ratio and its correlation with the 27Al/24Mg ratio inside a mineral sample, we can infer the amount of 26Al that was initially present in solid samples that formed at the time when the Sun was formed 4.6 Gyr ago (section 1.3). We can even measure the initial abundance of 26Al in stardust recovered in meteorites and produced around stars and supernovae that exploded before the formation of the Sun, between roughly 5 and 7 Gyr ago (section 1.4).

Due to these rich and far-reaching implications, the production and distribution of 26Al in the Galaxy has become the topic of many investigations (see, e.g. reviews by [46]). At the core of all such investigations are the nuclear reactions that produce and destroy 26Al inside stellar objects, from giant stars to novae and supernovae. Large uncertainties, for example, in the processes of mixing in stars and supernovae, and transport of radionuclides in the interstellar medium hamper the interpretation of the observational constraints. Nevertheless, we need to produce stellar yields that do not include significant nuclear physics uncertainties. This requirement is timely because models of supernovae, the Galaxy, and molecular clouds are improving rapidly and the description of various physics processes is becoming more detailed than ever. If the stellar yields are systematically incorrect, due to inaccurate reaction rates, then even the most sophisticated stellar and galactic models will provide us with the incorrect answers. Furthermore, in the case of stardust grains (presented in section 1.4), an almost direct signature of nuclear reactions is recorded in the grains and therefore reaction rates are essential to make any meaningful comparison between models and observations (see, e.g. [7]).

The aim of this review is to provide an updated, broad overview of the production and destruction of 26Al in different stellar objects and of the status of the reactions involved, in order to support stellar modellers with stronger and clearer information about the nuclear physics inputs to include in their calculations. Using stellar yields from these models we can then interpret observational data, either directly (as in the case of stardust grains, section 1.4) or by feeding the yields into models of the Galaxy, of the interstellar medium, and of Giant Molecular clouds. These allow us to use the abundance of 26Al as a tracer to address currently debated topics, from the physics of the interstellar medium to the circumstances of birth of the Sun.

Figure 1 shows the main nuclear reactions that directly affect the production and destruction of 26Al, which include both proton and neutron captures. These are relevant for most of the astrophysical sites responsible for the production of 26Al in the Galaxy and will be covered in detail here. There are many more reactions that affect the production and destruction of 26Al indirectly. For example, neutron source reactions such as the 22Ne(α, n)25Mg reaction [8] can affect the destruction of 26Al via neutron reactions in low-mass (section 2.1) and massive stars (section 2.2.2), while the bypass reaction 25Al(p, γ)26Si affects the production of 26Al in novae (see section 2.3). Iliadis et al [9] also listed the 25Mg(α, n)28Si, 24Mg(n, γ)25Mg, and 23Na(α, p)26Mg reactions as relevant, some of these will be discussed in section 3.5. In particular, the specific case of nucleosynthesis in core-collapse supernovae (CCSNe) (section 2.2.2) involves several different types of burning episodes and related reactions. A fully comprehensive analysis needs to be performed within the context of stellar physics uncertainties and is beyond the scope of this work.

Figure 1.

Figure 1. Section of the nuclide chart illustrating the main nuclear reactions favouring (green arrow) or inhibiting (red arrows) the production of 26Al in stellar objects. For clarity, we do not show the decay of the ground state of 26Al into 26Mg.

Standard image High-resolution image

In the sections below, we present an overview of the observational opportunities and implications related to 26Al, from γ-ray observations and Earth samples, to meteoritic stardust, and the early Solar System, and we finish the Introduction with some general considerations about the evolution of 26Al in the Galaxy. Section 2 describes more specifically the production of 26Al in different astrophysical sites, from asymptotic giant branch (AGB) stars (i.e. stars with initial masses roughly less than 10 M), to massive stars (i.e. with initial masses roughly above 10 M and and their CCSNe, and novae (i.e. accreting white dwarfs). For further reading on the topic of observations and models we refer to the review from an ISSI workshop series by Diehl et al [10]. Section 3 covers the reaction rates responsible for the production and destruction of 26Al in stars, from those directly involved with 26Al, to a selection of the most important indirectly related. Section 5 summarises the information presented here and proposes the future work needed to overcome the current major problems and uncertainties in the investigation of the production of 26Al in stars.

1.1. Live 26Al from γ-ray observations

The observation of cosmic radioisotopes relies on radioactive decay occurring outside the radioisotope stellar production sites, therefore, they are not distorted from absorption of photons by gas, which occur within the high density stellar matter. Hence, such astronomical data from radioactive decay convey direct information about nuclear reactions within cosmic sites that is otherwise hidden from direct observation (with the exception of neutrinos, which are, however, not observable from distant sites). Commonly available astronomical abundance data from atomic-line spectroscopy are also interpreted in terms of cosmic nucleosynthesis; this is, however, quite indirect information, in particular because the density and ionization state of atoms at the surface of stars is controlled not only by nuclear reactions but also by atomic processes that strongly affect the abundances. Therefore, the characteristic γ-ray lines measured from radioactive decay provide more direct astronomical data. The detection of the characteristic γ rays from 26Al decay is the first direct, convincing proof that nucleosynthesis is going on within the current Galaxy 30 , because 26Al has a characteristic decay time of a million years, much shorter than the age of the Galaxy of more than 10 billion years. Therefore, 26Al γ rays can be used to study recent nucleosynthesis sources and the transport of ejected material into the interstellar medium.

Direct observations of 26Al decay in interstellar space through its characteristic γ-rays with energy 1808.65 keV had been motivated by theorists [12], and was first achieved by the HEAO-C satellite in 1978/1979 observing the central regions of our Milky Way Galaxy [13]. The NASA Compton Gamma-Ray Observatory (1991–2000) with the COMPTEL instrument provided a sky image of the 26Al γ-ray line, which showed a structured 26Al emission, extended along the plane of the Galaxy [1416] (figure 2). This image was obtained from measurements taken over years 1991–2000 throughout the sky. It uses a maximum-entropy regularization together with the likelihood of the image's projected data fitting the measurement, and varies the image iteratively until a best image is found to fit the measured events. Such forward convolution analysis is required, as the direct inversion of measured data is not possible; the signal of the 26Al sky is translated into the data space of measured events by the instrument response function, which includes Compton scattering (as a probabilistic process), and is singular (hence cannot be inverted). The high instrumental background needs careful modelling, and Poissonian statistics must be properly included, hence the maximum-likelihood method is used (see [17, 18] for details on COMPTEL data analysis). The reliability of such γ-ray imaging has been consolidated in many studies (see, e.g. [19]). The 26Al γ-ray image and the structures it showed was found to be in broad agreement with earlier expectations of 26Al being produced throughout the Galaxy mostly from massive stars and their CCSNe since young stars are typically located on the plane of the Galaxy, and these young stars are preferentially massive.

Figure 2.

Figure 2. The 26Al sky as imaged with data from the COMPTEL telescope on NASA's Compton Gamma-Ray Observatory [14]. Reproduced with permission from [12].

Standard image High-resolution image

In 2002, the continuing INTEGRAL mission of the European Space Agency (ESA) started, with its high-resolution γ-ray spectrometer SPI, deepening the astronomical harvests of 26Al emission (figure 3). Note that the COMPTEL scintillation detectors had an instrumental resolution of ∼200 keV, compared to ∼3 keV for the SPI Ge detectors. This led to deeper, Galaxy-wide investigations of 26Al [20], as recently reviewed [21]. Furthermore, INTEGRAL allowed to spatially resolve specific and well-constrained massive star groups (OB associations), and therefore to test our understanding of massive star groups and how they shape the star-forming interstellar medium (see [22] for a review of astrophysical issues and lessons). Important herein are the Cygnus, the Orion [23] and the Scorpius–Centaurus [24] stellar groups. Altogether, astronomical 26Al observations have led to both the tracing of the path of nucleosynthesis ejecta after they leave their sources and eventually end up in next generation stars, and the investigation of the cosmic production sites, i.e. stars and supernovae. For the latter objective, it is essential to have the best-possible knowledge of the physics of the stellar sites and of the nuclear reactions involving 26Al, as discussed here.

Figure 3.

Figure 3. The 26Al line as seen with INTEGRAL high-resolution spectrometer SPI integrated on 13 years of measurements [25]. Reproduced with permission from [20].

Standard image High-resolution image

1.2. Live 26Al in terrestrial archives

Ice cores, deep-sea sediments, and deep-sea FeMn crust material are favourable locations to search for live radionuclides on Earth produced by nearby (in time and space) stellar nucleosynthetic events. These materials have very low growth rates, of the order of mm to cm per thousands to millions of years, and therefore they can provide time-resolved information over time scales of million years. However, because the number of radioactive atoms to be counted is tiny (e.g. 26Al is typically 12 to 16 orders-of-magnitude lower in abundance than the stable terrestrial 27Al), only accelerator mass spectrometry (AMS) [3] has so far reached the sensitivity required for such studies. AMS directly counts the radionuclide of interest one by one by means of a particle detector after the sample material is dissolved and the radionuclide of interest is chemically separated from the bulk material. Thanks to this methodology it has been possible to detect on Earth 60Fe from one or several nearby CCSNe that occurred roughly 2–8 Myr ago [2628], where 60Fe is another radioisotope with a half-life T1/2 = 2.62 Myr produced by massive stars and their supernovae, also observable via γ-ray satellites [29].

Detection of 60Fe of stellar origin is possible because its terrestrial production is negligible. In contrast, 26Al is produced not only in stars but also in the terrestrial atmosphere through cosmic-ray induced nuclear spallation reactions on abundant stable isotopes, such as 27Al(p, p–n), 26Mg(p, n), 28Si(p, 3He), and 24Mg(3He, p). Furthermore, production inside the terrestrial archives themselves and influx of interplanetary dust grains may add spurious amounts of cosmic-ray produced 26Al. Stellar 26Al may be of the order of only a few percent, up to roughly 10%, relative to its terrestrial component [3]. Therefore, to identify a stellar 26Al signal above the terrestrial background requires an extremely sensitive and efficient detection technique. Feige et al [3] searched for presence of stellar 26Al in an extensive set of deep-sea sediment samples that covered a time-period between 1.7 and 3.2 Myr ago and found an exponential decline of 26Al with the age of the samples that can be explained by radioactive decay of terrestrial 26Al. This indicates no significant 26Al above the terrestrial signal. Nevertheless, owing to the large number of samples analyzed, these data allowed the deduction of an upper limit for the stellar 26Al.

This upper limit is crucial when taken together with the previously derived stellar 60Fe, as the live 26Al and 60Fe radioisotopes found on Earth originated from nucleosynthesis in massive stars and, in particular, constrain models of the nearby CCSN that occurred 2 Myr ago. These are the same sources that dominate both the current abundances of live 26Al and 60Fe in the Galaxy and their extinct abundances in the early Solar System—since massive stars are present in star-forming regions (sections 1.1 and 1.3). Therefore, comparison between three different types of constraints can provide us with significant information on both massive star nucleosynthesis and the environment of the formation of the the Sun. The Earth samples produced a lower limit 60Fe/26Al between 0.1 and 0.33, which is close to the value obtained from Galaxy-wide gamma-ray spectrometry with INTEGRAL of 0.2–0.4 [29]. However, it is well above the 60Fe/26Al ratio derived for the early Solar System of 0.002. This calls for a different origin of 26Al in the early Solar System, relative to the current live 26Al in the Galaxy and in Earth archives. The problem is further compounded by theoretical CCSN yields, which currently overproduce 60Fe relative to 26Al, even compared to the spectroscopic γ-ray observations [3032].

1.3. Extinct 26Al in the early Solar System

Analysis of the isotopic composition of Mg in Ca–Al-rich inclusions (CAIs) from primitive meteorites, the oldest solid objects to have formed within the Solar System, demonstrates that 26Al was present in the early Solar System with 26Al/27Al ≃ 5 × 10−5. The isochrone method with which this is achieved is explained using figure 4, which presents the first observational evidence of the presence of 26Al. Several different minerals were analysed within the same sample (of same color in the plot), and for each of these minerals the ratios 27Al/24Mg and 26Mg/24Mg were measured and plotted against each other as in the figure. 27Al is the only stable isotope of Al, and 24Mg the most abundant stable isotope of Mg. Therefore, the 27Al/24Mg ratio is a good measure of how much total Al and Mg are present inside each mineral 31 . Because 26Mg is the daughter nucleus of 26Al, the 26Mg/24Mg ratio can trace the potential initial presence of 26Al: i.e. 26Mg/24Mg = 26Mg/24Mg + 26Mginitial/24Mg, where 26Mg is the radiogenic abundance from the decay of 26Al, and 26Mginitial its initial abundance in the material from which the mineral comes from, which is a constant. The discovery of the presence of 26Al is based on the fact that the 26Mg/24Mg ratios measured in different types of Mg-rich (i.e. low Al/Mg) and Al-rich (i.e. high Al.Mg) show a linear correlation with Al/Mg. The only way to explain such linear correlation is that the excess of 26Mg was initially due to an excess of 26Al. The slope of the line provide us with the initial 26Al/27Al ratio, given that the first term in the equation above can be written as 26Al/27Al × 27Al/24Mg (figure 4).

Figure 4.

Figure 4. Isochrone (solid line) obtained from the pioneering analytical work of [33], and later confirmed by [34]. The slope of this line allowed to obtain for the first time the canonical early Solar System 26Al/27Al ratio of ≃5 × 10−5 from specific CAI inclusions from different meteorites (Allende, Efremovka, and Murchison, different symbols). The 26Mg/24Mg ratio measured in the different minerals that make up the inclusions clearly correlates with the 27Al/24Mg ratio of each mineral. This proves that the excess 26Mg was built inside each mineral from the 26Al decay after, rather than before, the mineral formed. In the same words of the title of [33]: 26Al is a 'fuel' and not a 'fossil' in the early Solar System, i.e. it was present 'live', not already decayed. Note that there are also some samples that show no evidence of 26Al instead (open symbols). Data from [35], and references therein.

Standard image High-resolution image

The presence of 26Al in the early Solar System was actually predicted in the 1950s before its discovery, because an early heating source was needed to melt the interiors of the first planetesimals that formed within the first Myr [36]. This heat was driven by the γ-ray photons generated when 26Al decays to 26Mg inside a rocky body, the same photons that are observed via the INTEGRAL satellite when 26Al decays in the interstellar medium (figure 3). One of the consequences of such heat is that the ice inside those planetesimals that formed beyond the ice line (initially made of roughly 50% ice) melted and the water was lost. The timescale of this water loss was shorter than the accretion timescale of planetesimals into terrestrial planets. As sketched in figure 5 from [37], if 26Al is present in planetesimals, water-poor planets, such as the Earth, which is only a few percent water, result and are known to be habitable world. If 26Al is not present in significant quantity, instead, water-rich, ocean planets are predicted, which are potentially more difficult to harbour life. Therefore, the presence of 26Al in star-forming regions can strongly affect the habitability of the planetary systems formed in such regions. Understanding the production of 26Al in stars is therefore crucial to understand the water context of extrasolar terrestrial planets. However, the origin of 26Al in the early Solar System is still unclear and strongly debated with several separate stellar origin scenarios proposed 32 , each of which have different probabilities to occur in star-forming regions [6].

Figure 5.

Figure 5. Qualitative sketch of the effect of the 26Al radioactive fuel in the early Solar System, within the framework where water is carried to terrestrial planets from planetesimals that formed beyond the snow line. The left and right sides show the build up of terrestrial planets in potential 26Al-poor and 26Al-rich (like the Solar System) planetary systems, respectively. The middle grey arrow indicate the process of the accretion of planetesimal (the first sizable rocks with radius of the order of 50 km) that lead to the formation of rocky planets like the Earth. The blue and red arrows on the bottom right indicate the evolution of the planetesimal water content and of the abundance of live 26Al, respectively, for the case of 26Al-rich systems. The difference in the planets radius (ΔRP) between the two cases indicated at the top of the figure is a measurable quantity for extrasolar planets. Reprinted by permission from Springer Nature: Springer Nature Astronomy, 'A water budget dichotomy of rocky protoplanets from 26Al-heating', Lichtenberg, T; Golabek, GJ; Burn, R; Meyer, MR; Alibert, Y; Gerya, TV; Mordasini, C, 3, 307–313, 2019. Reproduced from [38], with permission from Springer Nature.

Standard image High-resolution image

For example, Cameron and Truran [39] were the first to propose a simultaneous enrichment plus triggered collapse scenario, in which a nearby CCSN ejects freshly synthesised material into a dense molecular cloud core thereby triggering its collapse to form the solar nebula (see e.g. [4042]). Alternate mechanisms for the injection of 26Al-rich material from a CCSN have been proposed, including, for example, pollution of the already formed disk [4345]. Other scenarios postulate that the gas that later collapsed to form the protosolar molecular cloud was pre-enriched in 26Al by an earlier generation of massive stars either within the star-forming region or the galactic interstellar medium itself [4651]. The winds of stars of mass > 30 M could also produce enough 26Al to be a candidate source [52, 53]. Brinkman et al [54] showed that lower mass stars could also achieve high yields if they are part of a binary system (section 2.2.1). Another possible origin is an AGB star [55], however, this source seems unlikely because these stars cannot produce the required 26Al abundance without also overproducing several radionuclides heavier than iron [35]. Furthermore, low-mass stars formed in the same molecular cloud as the Sun take a long time to reach the AGB phase (∼1 Gyr), by which time the molecular cloud is since long dispersed. Finally, that a molecular cloud would be visited by an AGB star formed elsewhere has been shown observationally to be very unlikely [56].

In summary, investigation of the production of 26Al by nuclear reactions in massive stars, and its ejection from massive stars, both during the wind phases and the following CCSN explosion, is relevant to understand the evolution and properties of planetesimals and planets both within solar and extrasolar systems.

1.4. Extinct 26Al in stardust grains

Stardust grains are recovered from meteorites and carry the pure signature of the nucleosynthesis that occurred in their parent stars, therefore, they are effectively tiny specks of stars [57, 58]. Evidence that 26Al was incorporated into stardust grains that formed around stars and supernovae is inferred from the Mg composition of the grains, and in particular from the excess in the daughter nucleus 26Mg, relative to the other stable Mg isotopes. Most of the grains originated from AGB stars, with some (a few percent) of grains also showing the signature of an origin from CCSNe and novae. Therefore, even if AGB stars and novae are not the major producers or 26Al in the Galaxy, relative to massive stars, it is necessary to investigate their 26Al production to interpret the presence of 26Al in stardust grains, at the time of their formation.

Many recovered and analysed stardust minerals are rich in Al and poor in Mg, so that the signature of 26Al is evident and measurable. These include silicon carbide (SiC) and graphite, which are carbon (C)-rich grains that form in a gas when C > O, and corundum (Al2O3) and hibonite (CaAl12O19), which are oxygen (O)-rich grains that form when C < O. If there are only just traces of Mg originally in the grains, then the full initial abundance of 26Al at the time of the grain formation can be recovered as the whole abundance of 26Mg. In other words, the abundance of 26Al is derived directly from the abundance of 26Mg because Mg is not a main component of the material: the Mg abundance is orders-of-magnitude smaller than that of Al in, e.g. aluminium oxides and silicon carbide grains (see, e.g. figure 2 of [59]), therefore, stable 26Mg is an orders-of-magnitude less significant contributor to the atoms at mass 26. 33 Groopman et al [60] significantly improved the derivation of the initial 26Al abundance by using the isochrone method for stardust grains as done for Solar System materials (section 1.3). This method produces more accurate results, generally showing higher ratios than previously estimated due to a better estimate of contamination.

Overall, C-rich stardust grains believed to have originated from CCSNe show very high abundances of 26Al, with inferred 26Al/27Al ratios in the range 0.1–1, higher than standard theoretical predictions in the C > O regions of the ejecta. These ratios can be used to constrain the nucleosynthesis models, and they require an extra production mechanism for 26Al to be at work in the C-rich regions of CCSNe beyond those described in section 2.2.2. This mechanism may be related to ingestion of hydrogen into the He-burning shell and the subsequent explosive nucleosynthesis [62]. The grains that are known to have originated in AGB stars show somewhat lower 26Al abundances than grains from CCSNe, with 26Al/27Al ratios in the range 10−3 to 10−2. Also these grains can be used to constrain AGB nucleosynthesis models [7, 63, 64]. For example, oxide grains belonging to a specific group (Group 2) show strong depletion in 18O/16O and have also relatively high 26Al/27Al ratios. Both features are a product of efficient hydrogen burning, possibly connected to hot bottom burning (HBB) in massive AGB stars [7] or extra-mixing in low-mass AGB stars [65] (see section 2.1). Finally, some grains have been potentially interpreted as nova condensates and their relatively high 26Al/27Al ratios, up to 0.3–0.5, have been used as one of the main indications of such signature origin [66, 67] (see section 2.3 for more details).

1.5. Evolution of 26Al in the Galaxy

The evolution of the average abundance of a radioactive nucleus such as 26Al in the galactic interstellar medium is generally controlled by the establishment of a steady-state equilibrium between its ejection by stellar sources and its radioactive decay. When such equilibrium is established, the number of 26Al nuclei N26 no longer changes with time t: ${{\rm{d}}{N}}_{26}/{\rm{d}}{t}=0$. The ${{\rm{d}}{N}}_{26}/{\rm{d}}{t}=0$ rate is made of two terms: one is a positive production term given by the stellar production rate per unit time, defined as dP26/dt; the other is a negative destruction term, wrought by the decay and equal to N26 × λ26 = N26/τ26, where λ26 is the decay rate of 26Al, and τ26 its mean lifetime. If the total abundance change is zero, then the two terms are equal and the equilibrium abundance N26 is equal to (dP26/dt)τ26. This formula, together with the stellar production rates predicted for the main stellar producers of 26Al in the Galaxy (see section 2), enables us to quickly compare the predicted 26Al abundance in the Galaxy to that inferred from γ-ray observations, which represents the sum of the contribution of all nearby young stellar populations currently ejecting this radioisotope.

However, this simple steady-state formula is not enough to interpret the data accurately. Theoretical models must also account for the star formation rate of the Milky Way today; the initial mass function, in order to model stellar populations; and the fact that stellar yields may vary with stellar mass and metallicity, as well as potentially their binary status. Furthermore, stellar enrichment within the interstellar medium is not continuous but represented by events discrete in time and space, therefore, it cannot be accurately described by a continuous dP26/dt production rate. Local enhancements or reductions in the 26Al abundance due to spatial and temporal inhomogeneities must also be considered in numerical simulations [68, 69]. This is a challenging task because the evolution of 26Al depends both on the time interval between the formation of the progenitor star that led to the enrichment event, a parameter that is currently poorly constrained, as well as their spatial distribution. Comparing 3D hydrodynamic simulated distributions of 26Al to the observed 1.8 MeV emission line flux maps of figure 2 [70, 71] has provided constraints on the Galaxy-wide distribution of 26Al, and showed that our observer position may be highly biased by the local environment and the stars that populated it. Furthermore, dedicated models [72, 73] are needed when considering specific star-forming regions and the 'super-bubbles' generated by the energy from massive stars within them.

Interpreting the abundances of 26Al derived from meteoritic data also requires the use of chemical evolution models, in this case specifically because the initial 26Al abundance in the mineral can be measured only relative to that of the stable 27Al. For the early Solar System, while the abundance of 26Al reflects the local star formation rate at the time of the formation of the Sun, the abundance of the stable isotope 27Al encodes the complete past enrichment history of the Galaxy and can only be predicted using galaxy models that properly integrate the galactic star formation history with stellar yields [7477]. These predictions are, however, affected by many uncertainties such as the total stellar mass formed prior to the formation of the Solar System, the amount of stable 27Al locked in stellar remnants (white dwarfs, black holes, neutron stars), and large-scale outflows that remove some of the 27Al content from the Galaxy [76, 77].

All the uncertainties described above can be significantly reduced by considering the 60Fe/26Al ratio, since 60Fe is a radioisotope with a similar half-life (2.62 Myr) and, like 26Al, can be measured via γ-rays and meteoritic analysis, and is also produced mostly by massive stars. Therefore, its ratio relative to 26Al solves some of the problems listed above, i.e. a chemical evolution simulation that follows the complete history of the Galaxy is not necessary to give us a direct insight into the stellar sources of these radioisotope. The steady-state equilibrium number ratio of the two radioisotope is N60/N26 = (dP60/dP26)(τ60/τ26), and the number ratio of the two radioisotope derived from the flux (F) ratio observed via γ-rays is N60/N26 = (F60/F26)(τ60/τ26) 34 . Therefore, the 60Fe/26Al production ratio in massive stars (by number) can be directly compared to the observed flux ratio [29]. The result is that current models overproduce 60Fe relative to 26Al compared to the spectroscopic γ-ray observations by a factor of 3–10 [3032], and there are inconsistencies not only between models and observations, but also between observations of different types (see end of section 1.2). Moreover, in reality, local inhomogeneities can also affect the 60Fe/26Al ratio, even if these two radioisotope came from exactly the same sources, since they decay with different half-lives. Considering temporal heterogeneities within a statistical framework results in an increase of roughly 10% in the 60Fe/26Al ratio relative to that calculated using the basic steady-state formula [78], but even larger fluctuations are found in more sophisticated models of the interstellar medium [51] and of giant molecular clouds [79].

Finally, we stress that the fundamental input for any numerical predictions of the evolution of 26Al in the Galaxy—whether associated with chemical evolution models, interstellar medium simulations, or single-source enrichment models are the adopted stellar yields discussed below and e.g. in [9, 8083]. In fact, the quantitative predictive capabilities of even the most sophisticated simulations are still limited by the large uncertainties affecting nucleosynthesis and stellar evolution calculations. In the following section, we describe in more detail the production of 26Al in different types of astronomical objects, from low to high mass stars and binary interaction objects, and the associated uncertainties. Lower mass stars during their AGB phase (section 2.1) are specifically relevant to understand the origin of 26Al in the majority of meteoritic stardust grains. Massive stars and their CCSNe (section 2.2) are the dominant producers of 26Al in the Galaxy and in star-forming regions, which are relevant for the early Solar System. Novae are also relevant as potential contributors to galactic 26Al, as well as sources of rare stardust grains; they will be discussed in section 2.3.

2. Stellar production sites

2.1. Asymptotic giant branch (AGB) stars: extra-mixing and HBB

Stars with initial mass in the range 1–8 M end their lives as AGB stars. Their structure consists of a degenerate carbon–oxygen core, above which sit a helium-burning shell and a hydrogen-burning shell separated by an He-rich region called the 'intershell'. Surrounding this central region of the star is an extended, hydrogen-rich, convective envelope. The configuration of thin helium and hydrogen burning shells is unstable, leading to periodic runaway helium-burning episodes, known as thermal pulses. During these thermal pulses, convection mixes the ashes of helium-burning in the intershell region. As a thermal pulse subsides, the star undergoes a structural readjustment that allows the convective envelope to deepen, penetrating the intershell and dragging freshly synthesised material to the surface. This is known as the third dredge-up and it may happen after each thermal pulse. Depending on the strength of the stellar winds (which strip mass from the surface) and the mass of the envelope, an AGB star may undergo a few to many tens of such thermal pulses before the entire envelope is removed and the star transitions to the post-AGB phase. For detailed reviews of AGB stars and their evolution, we refer the reader to Herwig [84] and Karakas and Lattanzio [85].

The production of 26Al in AGB stars has been studied by many authors [35, 63, 86, 87]. Both the hydrogen-burning and the helium-burning shells are relevant to synthesis of 26Al. In the latter, the initial abundances of the two heavy magnesium isotopes, 25Mg and 26Mg, are enhanced via α-captures onto 22Ne, and then dredged-up into the stellar envelope. Directly at the base of the envelope, or in the radiative region just below it, depending on the initial stellar mass, shell hydrogen burning converts 25Mg into 26Al via proton captures as part of the Mg–Al chain shown in figure 6.

Figure 6.

Figure 6. Main proton-capture reactions and β+ decays (with half-lives indicated) involved in the MgAl cycles. The solid- and dashed-line boxes represent stable and unstable nuclei, respectively. The red arrows show the reactions and β+ decays that result in the by-passing of the production of the ground state of 26Al, due to the production of the isomeric state instead. In particular the by-pass via 26Si can be activated at high temperatures (200–400 MK) during the nova nucleosynthesis described in section 2.3, while the by-pass via 26Mg is also activated at the relatively low temperatures of AGB stars (from roughly 60 to 100 MK).

Standard image High-resolution image

Above initial masses around 4 M, the exact value depending on the metallicity and the choice of the convective model, the base of the stellar convective envelope is deep enough that it lies within the upper regions of the hydrogen-burning shell. Convection therefore cycles material from the entire envelope of the star through the hydrogen burning shell, in a process referred to as HBB. The combined action of the third dredge-up adding freshly synthesised 25Mg to the envelope, and the HBB processing this material via the Mg–Al chain makes massive AGB stars substantial producers of 26Al [87, 88]. This is in agreement with the high 26Al/27Al ratios in Group 2 oxide grains, and make massive AGB stars that experience HBB a candidate source of these grains [7], together with the Cool Bottom Process in low-mass stars are discussed below. The effect of reaction rate uncertainties on HBB in massive AGB stars has been carefully examined by Izzard et al (2007) [89], who concluded that uncertainties in the 25Mg(p, γ)26Al and 26Al(p, γ)27Si rates lead to an uncertainty in the 26Al yields from a factor of few at solar metallicity to up to two orders-of-magnitude at lower metallicities. The so-called super-AGB stars also evolve to the AGB but experience carbon burning in the core after the helium-burning has taken place. This results in ONe- rather than CO-rich cores 35 . The initial masses of these stars are in the range 8–10 M, depending on metallicity. In these super-AGB stars the base of the convective envelopes is very hot (up to 100 MK) and therefore HBB is very efficient leading to significant production of 26Al. Nevertheless, the overall super-AGB contribution to the estimated galactic content of roughly 2.8 M does not exceed 0.3 M, i.e. 10% of the total [90].

In canonical low-mass AGB models, where only convective mixing is taken into account, the hydrogen-burning shell is separated from the convective envelope by a radiative 'buffer' region where no mixing occurs. Therefore, the 26Al produced by H burning in the top layers of the H-burning ashes 36 is ingested inside the thermal pulse, where it can be destroyed by neutron-captures with neutrons generated by the 22Ne(α, n)25Mg reaction. What is left is then carried to the stellar surface via the third dredge-up. By this mechanism, the surfaces of low-mass AGB stars are predicted to have 26Al/27Al of the order of a few 10−3, in qualitative agreement with those observed in the SiC grains that originated in these stars [63].

The main stellar model uncertainties associated with AGB stars that impact the production of 26Al are the efficiency of the third dredge-up, the choice of mass-loss rate, and the convective mixing model. The efficiency of the third dredge-up is affected by the method employed to find the convective boundary, e.g. [91]. More efficient third dredge-up in massive AGB stars, for example, carries more 25Mg from the intershell into the envelope, which is then processed into 26Al via HBB. The mass-loss rate influences the duration of the AGB phase, with stronger mass loss leading to fewer thermal pulses and less nucleosynthesis [92]. The choice of the convective model affects the temperature structure of the stellar envelope and therefore the temperature at which HBB can take place. More efficient convection leads to higher HBB temperatures, favouring 26Al production [93, 94]. It can also lead to higher stellar luminosities, which may accelerate mass loss [93].

Another main uncertainty is related to the possible occurrence of non-convective mixing in AGB stars with initial mass below 2–3 M. This mixing may allow the crossing of the radiative buffer region and carry ashes from the hydrogen-burning shell into the convective envelope during the periods in-between thermal pulses (refereed to as interpulse periods) and boost the production of 26Al. Wasserburg et al [95] first suggested a non-convective mixing process, which they called Cool Bottom Process (CBP, in contrast to HBB in the most massive stars) as an explanation for some CNO isotope anomalies observed in low-mass AGB stars and stardust grains of AGB origin. In the CBP model, it is assumed that material is carried from the lower edge of the convective envelope down to the innermost layers of the hydrogen shell where it experiences proton-capture reactions and then returns it to the convective zone.

The work of [96] showed that the surface of AGB stars with mass ≤2 M and close to solar metallicity can be enriched in 26Al up to 26Al/27Al = 0.1. However, they did not provide any hypothesis on the physical mechanism driving the mixing, and treated the depth and the mixing rates as free parameters.

Furthermore, to reach 26Al/27Al = 0.1 in the stellar envelope the CBP model of the quoted authors must push the carried materials down to the deepest layers of the H-shell, where the temperature is greater than 5.5 × 107 K, before returning to the stellar envelope. The so-induced circulation of hot (still burning) matter can strongly affect the stellar energy balance, with relevant luminosity feedback.

In the last two decades many studies have been carried out on non-convective mixing phenomena of low-mass red giants and numerous hypotheses have been formulated on their cause: from stellar rotation [97], to thermohaline mixing [98], and gravity waves [99], stellar magnetic fields [100], and their combined effect [101, 102]. In the case of 26Al the problem remains that to synthesize this radioisotope in significant amounts via extra-mixing, material must experience relatively high temperatures, and most of the proposed mixing models listed above do not predict 26Al production. For example, the average molecular weight inversion due to the 3He(3He, pp)4He reaction, which triggers the thermohaline mixing, occurs in the H shell where the temperature is ∼3.5 × 107 K, too low to efficiently activate the 25Mg(p, γ)26Al reaction [64, 103]. Only the extra-mixing induced by the effects of the stellar magnetic field proposed in [104] is currently able to produce the 26Al/27Al ratios of up to a few 10−1 measured in oxide stardust grains of Group 2, also showing strong deficits in 18O, an isotope efficiently destroyed by H burning [105].

Magnetic-induced mixing in very low-mass (≤1.5 M) AGB stars can predict high 26Al/27Al because it operates via small bubbles of magnetized material, which rise from the H-burning shell to the base of the convective envelope [65], instead of moving material from top to bottom as in the classic CBP model. With state-of-the-art methods, the magnetic extra-mixing predictions of the Al isotopic ratios in low-mass AGB stars do not appear to be significantly affected by reaction rate uncertainties. For example, when using the rates by [106] and [107] for the (p, γ) captures on 25Mg and 26Alg, respectively, instead of those reported by [108] the changes in the resulting 26Al/27Al isotopic ratio is smaller than the variations due to the stellar model parameters (such as the stellar mass, the mass loss, and the mixing depth). However, the magnetically induced extra-mixing is relatively fast and the transported material has no time to be affected by proton captures along the path from above the H-burning shell to the envelope (as instead it might happen in the classic CBP). Therefore, any change induced by the mixing in the stellar surface composition reflects the abundances in the H-burning shell. As a consequence any uncertainty or change in the nuclear physics input that affects the 26Al/27Al distribution in the shell will also affect the one in the stellar envelope (see the brief analysis by [109]).

2.2. Massive star as the main source of 26Al in the Galaxy

From the map of the γ-ray observations from the 26Al decay (see, e.g. figure 2), it is clear that most 26Al is confined to the galactic plane and to some specific clumps (see, e.g. figure 16 of [110] and [111113]). These clumps coincide with known groups of massive stars (i.e, with initial mass > 10 M 37 ), usually referred to as 'OB associations', due to the fact that these massive stars are blue/blue-white stars with surface temperatures roughly above 20 000 K belonging to the spectroscopic O and B classes [114, 115]. The 26Al γ-ray map therefore indicates that most of the 26Al in our Galaxy is produced by massive stars [19, 116, 117]. Theoretical models in fact predict that massive stars eject large abundances of 26Al, both through stellar winds and CCSN explosions. Like in most of other stellar sources, 26Al is mainly made via the 25Mg(p, γ)26Al reaction, while there are two main destruction channels: proton and neutron capture reactions, depending on the burning phase. Here, we focus first in section 2.2.1 on the winds from both single massive stars and massive stars that have a companion in a close binary system. We also discuss very massive stars (VMS), i.e. with masses above 100 M [118]. Second, in section 2.2.2 we discuss in detail the contribution to 26Al coming from the CCSN explosive ejecta.

Overall, for stars of masses below roughly 40 M, the contribution of the explosive ejecta is typically dominant relative to that of the winds. For example, when looking at table 3 of Limongi and Chieffi [119], both contributions to 26Al, from the hydrostatic and explosive phases (see section 2.2.2), are ejected during the explosion. In stars of higher masses, instead the winds are stronger (see section 2.2.1) and therefore their contribution to the total amount ejected increases to become similar to that ejected during the explosion. Neutrino processes are not usually included in the CCSN models, but can contribute a relatively minor component of 26Al, as we present at the end of section 2.2.2. It should be kept in mind that, as discussed in detail below, these different contributions are strongly model-dependent. Stellar rotation and binarity, mixing in the hydrostatic and explosive phases of the stellar evolution, CCSN properties, such as the mass of the remnant and the energy of the explosion, as well as the choice of the reaction rates can all affect the different contributions to the total 26Al yield from massive stars of different masses.

2.2.1. Massive star winds in single and binary systems

The 26Al that is expelled from a massive star by winds is produced by hydrogen burning both in the core and, later, in a shell via proton-captures on 25Mg. This 26Al can be ejected by stellar winds when layers that once belonged to the H-burning core are exposed at the surface (in the case of Wolf–Rayet (WR) stars) or in the non-WR regime if some mixing mechanism allows the 26Al produced in H-burning regions to diffuse into the stellar envelope up to the surface. Convection, rotational mixing in the radiative zones of the star and mass losses are key processes that may enrich the surface in 26Al. For all massive stars, after the main sequence, the convective envelope penetrates into the layers where 26Al has been produced and transports this radioisotope to the stellar surface. The strong winds following the main sequence expel these layers from the stars, carrying the 26Al into the surrounding interstellar medium (ISM). For WR stars, the stellar winds already start during hydrogen burning and are so strong that the entire hydrogen-rich envelope is removed, all the way down to the top of the layer processed by the CNO-cycle (also known as the helium core), and this eventually includes the hydrogen-burning shell. The deeper layers exposed this way contain more 26Al, leading to increased wind yields for these stars. After core hydrogen burning, 26Al is quickly destroyed in the core during helium-burning by neutron capture reactions, specifically the (n, p) and (n, α) channels. However, because the convective helium-burning core is smaller than the original convective hydrogen-burning core, enough 26Al survives to be expelled from the star into the ISM.

At solar metallicity, Z = 0.014, all single massive stars above 30 M experience phases of strong mass loss during their lifetimes. The amount of 26Al ejected by the winds of massive stars, considering the currently recommended mass loss rate prescriptions (see [120] for the hot phase, [121, 122] for the cool phase, and [123] or [124] for the WR phase), depends in a sensitive way on the initial metallicity. At very low metallicity the amount of 26Al ejected is much smaller than at solar metallicity because the winds are expected to be much weaker, and therefore the mass loss is smaller. Figure 7 shows the 26Al yields ejected by the stellar winds as a function of initial stellar mass for models of solar metallicity predicted by various studies reported in the literature for non-rotating, solar metallicity stars. The highest yields are for stars with initial masses >30–40 M. These stars become the so-called WR stars mentioned before, as they lose their entire hydrogen-rich envelope, exposing their helium cores. During the phase where layers that belonged to the core H-burning phase (which are enriched in helium and nitrogen) are exposed at the surface, large amounts of 26Al are expelled into the ISM by these stars. Massive stars with masses between 10 and 30 M instead do not lose enough mass to expose the layers that have been processed by CNO burning.

Figure 7.

Figure 7. The 26Al yields (in the form of total mass of 26Al ejected by the winds in units of solar masses) as function of the initial stellar mass from various studies reported in the literature for non-rotating single stars at solar metallicity [81, 125128]. The red line gives the effective binary yields, i.e. the average increase of the yields of the primary star when considering a flat distribution for the binary periods and non-conservative mass-transfer, which assumes that all the mass transferred to the secondary star is subsequently ejected by the system and therefore results in an upper limit for the yields. The dotted pink line gives the VMS yields deduced from the models by [129] (figure based on figure 3 from [130]).

Standard image High-resolution image

Interestingly, most massive stars are found in binary systems (see e.g. [131]) and are close enough to each other to interact. Sana et al (2012) [132] found that more than 50% of all O-type stars (stars with initial masses from 15 M and higher) will interact with their companion during their lifetimes. More than 25% of these stars will interact with their companion even before the end of the main sequence [133]. These interactions influence both the evolution of the stars, and the 26Al yields [125, 134]. The process resulting from binary interaction that mostly influences the yields is the mass-transfer between the two stars due to Roche lobe overflow. The evolution of the primary (i.e. the initially heavier star of the binary) can be strongly affected by this, as compared to their single-star counterparts. For example, the time at which mass loss starts and the amount of mass lost by the star can be influenced by binary interaction. The main effect of the mass-transfer is that more mass is lost from the primary stars, uncovering the layers rich in 26Al that otherwise would not be uncovered if the star was evolving in isolation (for more details see [125]). The 26Al injected into the ISM by binary systems is a combination of the 26Al present in the layers stripped by mass-transfer and expelled, and the 26Al present in the deeper layers of the star that are driven off by the stellar winds (see figures 10 and 11 of [125]). The effect of binary interactions can be seen in figure 7 by comparing the red and the blue lines. The yields of the primary stars increase strongly below 30–40 M, up to a factor of 100 for the least massive stars (10–15). Above 30–40 M, the binary interactions do not increase the yields of the stars anymore, as the winds are already very efficient in ejecting 26Al.

Aside from massive star binaries and Wolf–Rayet stars, VMSs can also contribute significantly to the 26Al enrichment of the ISM. The opinion widely held until 2010 was that the most massive stars in the Universe are of an order of 100–150 M. More recent studies, however, have found evidence for the existence of VMSs of higher masses [135137]). These stars are thought to dominate the mechanical energy input and ionising radiation [138, 139] of star-forming regions such as the Tarantula nebula. Moreover, the mass-loss rates of VMSs have been shown theoretically [140] and observationally [136] to be larger than previously assumed via mass-loss rates of Vink et al 2001 [122]. The discovered upturn in mass loss for the most massive stars occurs at a transition point where optically thin O-star winds transform to becoming optically thick [140, 141]. Recent population synthesis models [142] including stellar evolution models for VMSs [129, 143] with the new higher mass-loss rates [140] need to be constructed to predict the likely dominant contribution of VMSs to the galactic 26Al emission budget. We expect such contributions to be high because VMSs are nearly homogeneous stars: their convective core during the main sequence phase extends over more than 90% of the total mass of the star. This means that almost all 25Mg initially present in the star can be converted into 26Al via proton captures and the reservoir of matter enriched in newly produced 26Al may correspond to almost the whole star. Furthermore, VMSs are very luminous and are expected to have strong radiation-driven winds. These two factors lead us to predict large amounts of 26Al to be ejected from these stars. For example, the non-rotating 500 M model by [129] shown in figure 7 loses about 450 M during the main sequence, leading to an 26Al yield of 7.6 × 10−3 M. This is a factor 100 more 26Al than for a 60 M star computed with the same code. Of course, these stars are very rare, however, their yields are so large that even a few events can have a strong impact on the galactic 26Al budget.

The main uncertainties that affect the production of 26Al in massive stars are of two types: those involving the nuclear reaction rates, in particular the 25Mg(p, γ)26Al rate and its branching ratio to the ground state of 26Al, and the 26Al(p, γ)27Si rate; and those involving the physics input of the stellar models, including: the size of the convective core, the mass-loss rates, either induced by radiative line driven winds or by mass loss due to mass-transfer episodes in close binaries, and the mixing processes in the radiative zones of the stars, such as turbulence induced mixing due to rotation. An increase either in either the size of the convective core, the mass-loss rates, or the mixing in radiative zones, would lead to an increase of the 26Al yields. See e.g. the discussion in Palacios et al [80].

2.2.2. CCSN from massive stars, hydrostatic and explosive nucleosynthesis

Several studies have been dedicated to the production of 26Al in massive stars including their following CCSN explosions [9, 75, 81, 144]. On the one hand, it is expected and shown by theoretical stellar simulations that the H-burning and C-burning ashes are 26Al-rich, as compared to the rest of the ejecta because these regions are rich in the protons needed to produce 26Al (protons are produced during C-burning via the 12C(12C, p)23Na reaction). On the other hand, during He-burning there are no protons available to efficiently produce 26Al, and if there is any 26Al already present here, it is typically destroyed by neutron captures, with neutrons produced by the 13C(α, n)16O and 22Ne(α, n)25Mg reactions. In more advanced stages like explosive O-burning and Si-burning, neither 25Mg or 26Al are efficiently produced [75].

While these general guidelines are derived from the nuclear astrophysics properties leading to the nucleosynthesis of 26Al, stellar models show significant variations due to intrinsic properties and simulation parameters. In figure 8 we show the 26Al mass fraction profile as function of mass coordinate for a M = 15 M star and a M = 20 M star model after the explosion [145], together with the profile of other major isotopes indicative of the stellar structure, and of 25Mg, the seed nucleus for the production of 26Al. Moving from the outer mass coordinate towards the center of the star, the first 26Al peak is observed in the H-burning layers for both models, this derives from destruction of most of the initial 25Mg via proton capture. Just below the H-burning layers, in the upper part of the H-poor, He-rich shell there is a significant amount of 26Al left from the burning that happened before the CCSN. For the 15 M model, for example, the peak is around a mass fraction of 2 × 10−5. This peak is due to either direct mixing of 26Al from the H-burning region just above, or to mixing of 14N, which provides protons via the 14N(n, p) 14C reaction, using the neutrons produced by 13C(α, n)16O. In the 20 M model this peak abundance is much more restricted in mass than in the 15 M model. At the bottom of the He shell, instead, 26Al is completely destroyed by the neutron captures triggered during the CCSN explosion in this region of the stars (the so-called 'neutron burst', or n-process, [146148]). Further differences between the 15 and the 20 M models are seen in the C-burning ashes. While the 26Al abundances before the explosion are comparable in the two cases (with a mass fraction of the order of few 10−6), the 20 M model shows a much stronger explosive production than the 15 M model, visible at a mass coordinate roughly 3.6 M, and comparable to the abundance made by H-burning.

Figure 8.

Figure 8. Final mass fraction distributions of the isotopes H, 4He, 12C, 16O and 28Si, indicative of the stellar structure, and of 26Al and its main seed isotope 25Mg as function of mass coordinate for the ejecta of CCSN models with initial mass M = 15 M (top panel) and M = 20 M (bottom panel) and initial metallicity Z = 0.02 [145].

Standard image High-resolution image

Alternative nucleosynthesis conditions from those described above may also be present and provide a significant contribution to the amount of 26Al ejected by a CCSN, for instance, these can be related to convective-reactive events during the hydrostatic evolution of the progenitor star. The work of [62] showed that a significant amount of 26Al can be also produced in the explosive He-burning ejecta, following the ingestion of H in the convective He-shell [149, 150].

Finally, neutrinos from the collapsing stellar core leading to the ν-process [151, 152] also affect the production of 26Al in two ways: directly via the 26Mg (νe , e) reaction, and indirectly by providing additional protons for the 25Mg (p, γ) reaction, mostly from 20Ne $({\nu }_{x},{\nu }_{x}^{\prime} p)$ and spallation from other abundant nuclei. The cross sections for the neutrino-induced reactions are very well constrained [153], in particular, the Gamow–Teller strength for 26Mg (νe , e) has been determined by charge-exchange reactions [154]. The contribution of this process to 26Al, however, is very sensitive to the rather uncertain neutrino-energy spectrum. Depending on the neutrino energies, the ν process may lead to an increase of the 26Al yields by 10%–40% [75, 153].

2.3. Nova outbursts

Classical novae are the stellar explosions that take place in stellar binary systems, consisting of a compact, white dwarf star and a low-mass companion, typically a K or M main sequence star, although observations increasingly reveal more evolved companions. Novae exhibit a sudden rise in optical brightness, with peak luminosity reaching 104–105 solar luminosity. During the explosion, roughly 10−5–10−3 M of material is ejected into the interstellar medium, at a speed of several 103 km s−1. Novae are expected to recur with typical periodicity between 1 and 100 year (recurrent novae) and 104–105 year (classical novae).

The main nuclear reactions involved in the production and destruction of 26Al in novae, and their associated uncertainties, have been discussed in several papers (e.g. [155, 156]) and are illustrated in figure 6). The synthesis of 26Al in novae requires the presence of some seed nuclei, such as 24,25Mg, or to some extent, 23Na, and 20,22Ne. Since novae do not achieve high enough temperatures to power CNO-breakout, 26Al production requires an underlying ONe white dwarf (rather than a CO white dwarf) and some mixing to occur at the interface between the outer layer of the core and the envelope. The main nuclear reaction path leading to 26Al is 24Mg(p, γ)25Al(β+)25Mg(p, γ)26gAl, whereas destruction is dominated by the 26gAl(p, γ)27Si reaction. The current main source of nuclear uncertainty comes from the rate of the 25Al(p, γ)26Si reaction: because 26Si only decays to the isomeric state of 26Al this reaction determines the fraction of the nuclear path that proceeds through the isomeric 26iAl state, via the decay of 26Si thus by-passing 26gAl synthesis.

Already before the discovery of 26Al in the interstellar medium by the HEAO-3 satellite through the detection of the 1809 keV γ-ray line [157, 158], Ward [159] suggested that 26Al could be produced efficiently in astrophysical environments such as classical nova outbursts, characterized by a rapid rise to maximum temperatures around Tpeak ∼ (2–3) × 108 K, followed by a relatively fast decline. One-zone, explosive H-burning nucleosynthesis studies corroborated this idea (see, e.g. [160, 161]), and concluded that while classical novae might produce sufficient amounts of 26Al to reproduce some of the observed isotopic anomalies found in meteorites, they would not represent major galactic factories of 26Al. These calculations, however, assumed solar composition (or CNO-only enhanced) envelopes. With the advent of models of nova explosions on ONeMg white dwarf stars [162], one-zone nova nucleosynthesis models predicted large amounts of radioisotope (such as 22Na and 26Al) in their ejecta [163, 164], suggesting that these novae might represent significant, though still not dominant, sources of the Galactic 26Al.

The following 1D hydrodynamic simulations [165, 166] stressed the crucial role played by convection in carrying a fraction of the fresh 26Al synthesized at the base of envelope to the outer, cooler layers where destruction through proton captures could be prevented. Still, the composition adopted for the underlying ONeMg white dwarfs adopted in these models was too crude as it was based on calculations of hydrostatic C-burning nucleosynthesis by Arnett and Truran [167] with mass fraction ratios X(16O):X(20Ne):X(24Mg) of 1.5:2.5:1. Stellar evolution models of intermediate-mass stars [168, 169] revealed that ONe white dwarfs are instead made basically of 16O and 20Ne, with above ratios of 10:6:1. The dramatic reduction in the 24Mg seeds resulted in a significant decrease in the contribution of novae to the galactic 26Al predicted by the first 1D hydrodynamic simulations from accretion to ejection for a realistic composition of the underlying white dwarf, and with updated nuclear reaction rates [170, 171].

Since the late 1990s, all hydrodynamic 1D nova simulations systematically resulted in some 26Al production. This includes the most recent 12321 nova models [172], which include the effect of the inverse energy cascade that characterizes turbulent convection in nova outbursts on the time-dependent amount of mass dredged-up from the outer white dwarf layers, and a time-dependent convective velocity profile throughout the envelope, as computed by 3D simulations [173175]. While these state-of-the-art models yield more massive envelopes than those previously reported, and result in more violent outbursts characterized by higher peak temperatures and greater ejected masses, their 26Al yields are similar to previous estimates for ONe novae (see [172] for details).

A crude estimate of the contribution of novae to the amount of 26Al present in our Galaxy can be obtained from [163, 170]:

Equation (1)

where Mejec is the mean ejected mass during a nova outburst, X(26Al) is the mean mass fraction of 26Al in the ejecta, f(ONe) is the fraction of ONe novae (typically, 1/3; see Livio and Truran 1994), Rnova is the nova rate ($\sim {50}_{-23}^{+31}$ year−1 [176]), and τ(26Al) is the mean lifetime of 26gAl (1 Myr). From these estimates, and adopting a relatively favorable ONe nova model (e.g. Mejec(26Al) ∼ 2 × 10−8 M [171]), we obtain an upper limit to the contribution of novae to the Galactic 26Al content of ≤ 0.34 M. This corresponds to about 12% of the Galactic 26Al (∼2.8 ± 0.8 M [177]) and is in qualitative agreement with the analysis of COMPTEL/CGRO 1.809 MeV 26Al emission map [177, 178], which favors younger progenitors (i.e. WR stars and CCSN, as discussed in section 2.2). However, it is worth noting that this estimate is affected by large uncertainties, since the mean ejected mass per nova outburst and the variation of the nova rate since the formation of our Galaxy are not well constrained.

3. Relevant nuclear reactions

26Al has two long-living states: the ground state with spin and parity Jπ = 5+ and half-life T1/2 = 0.717(24) My, and an isomeric state 26Alm at 228 keV with Jπ = 0+ and T1/2 = 6.3460(8) s [1]. It is the abundance of the 26Al ground state that is relevant for the cosmic 1809 keV γ-ray flux. Namely, the β+-decays from the ground state of 26Al feed the first excited state in 26Mg, which de-excites by emitting 1809 keV γ-rays. The isomeric state 26Alm , instead, decays via a fast superallowed β decay directly to the 0+ ground state of 26Mg and hence does not contribute to the cosmic γ-ray flux from 26Al.

In astrophysical environments (see section 2), 26Al is mainly produced via radiative proton captures on 25Mg (see section 3.1). There, de-excitations from the populated states feeding ultimately either the ground or the isomeric state in 26Al are relevant in order to determine the population ratio between the two states of 26Al. At lower temperatures, these two states act like separate nuclear species, but at temperatures above 0.4 GK, thermal excitation populating the isomeric state starts to play a role and reduce the effective lifetime of 26Al, as discussed in section 4. Moreover, at high stellar temperatures and densities, capture of free electrons from the continuum will also affect the lifetime of 26Al (see section 4). In addition to the production reaction 25Mg(p, γ)26Al, key reactions to determine the abundance of 26Al are the destruction p-capture reaction 26Al(p, γ)27Si (section 3.2), the neutron-induced reactions on 26Al (see section 3.3) and the bypass route via 26Si (see section 3.4). Other reactions relevant for the production and destruction of 26Al are briefly discussed in section 3.5. The reaction network around 26Al summarizing the relevant nuclear reactions are shown in figures 1 and 6. In summary, in table 7 all the reactions considered here are listed, together with their relevant stellar sites, and corresponding temperatures.

We have recalculated some of the reaction rates presented in the following subsections using the RatesMC code [179] based on the information presented in the tables in each subsection. The resonance energies have been recalculated from the excitation energies, the separation energies and including the corrections for atomic binding as in [180]. Details as to resonance strengths and/or proton widths are provided in each subsection, as needed, together with figures of the new rates. The tabulated data for the rates can be found in tables 812.

3.1. The production reaction 25Mg(p, γ)26Al

The rate of the 26Al production reaction is dominated by resonant proton captures to levels above the proton threshold at 6306.33(6) keV [181] in 26Al. Direct proton captures to bound states in 26Al and the contribution from the subthreshold (3+) resonance at ${E}_{{\rm{res}}}=-24.86(10)$ keV are negligible compared to the resonant captures at temperatures above ≈0.006 GK [182]. Therefore, we focus here on the resonant proton captures. Table 1 summarises the available data on the excitation energies, spins and parities as well as the resonance energies and strengths for the relevant states in 26Al. The ground state spin and parity of 25Mg is 5/2+, and therefore = 0 proton captures populate states with Jπ = 2+ or 3+ in 26Al, = 1 proton captures states with Jπ = 1, ..., 4, and = 2 states with Jπ = 0+, ..., 5+. All states, except the Jπ = (7) state at 6695 keV, can be populated via = 0 − 2 proton captures, however, many of the spin-parity assignments are still tentative (see table 1) and further studies are needed.

Table 1. Excitation energies (Ex ) together with the spins and parities (Jπ ) for the excited states above the proton separation energy (Sp = 6306.33(6) keV [181]) in 26Al. The excitation energies, spins and parities are from from [1] unless stated otherwise. The resonance energies have been recomputed from the excitation energies include a correction for atomic binding of ΔBe = − 1.14 keV. The resonance energies (Eres), and experimentally determined resonance strengths ω γ for the relevant states are given. The resonance strengths are the total resonance strengths—the f0 factors are the branching of the resonance to the ground state of 26Al. The f0 has been recomputed [183] from the listed values in the ENSDF repository including the full cascade to the ground state and the isomer. Only states up to Ex = 6800 keV are listed.

Ex (keV) Jπ Eres (keV) ω γ (eV) fa
6280.33(9)(3+)–24.85(11) ${\theta }_{p}^{2}=0.0127{(26)}^{0}$ 0.770(3)
6343.46(8)4+ [184]38.29(10)4.9(21) × 10−22 b [184]0.79(5)
6363.99(8)3+ 58.81(10)2.9(5) × 10−13 [185]0.81(5)
6398.64(21)2 [186]93.46(22)3.5(4) × 10−10 [187, 188]0.67(4) [188, 189]
6414.46(10)(0 to 2+)109.28(12)2.3(1) × 10−11 c [190]0.0042(6) d
6436.44(11)(3 to 5+)131.26(13) ≤ 2.5 × 10−10 [187]0.727(3)
6495.94(7)(3 to 5+)190.76(9)5.2(36) × 10−7 e 0.75(2) [187]
6550.68(7)(4+, 5)245.50(9)55.0(63) × 10−7 [179]0.80(1)
6598.32(16)(5+)293.14(17)45.0(52) × 10−6 [179]0.71(1)
6610.40(6)(3)305.22(8)2.8(3) × 10−2 f 0.878(10) g
6680.45(7)(2+)375.27(9)6.0(6) × 10−2 [191]0.67(1)
6724.25(7)(4)419.07(9)7.4(2) × 10−2 [191]0.96(1)
6783.79(5)(2)478.61(8)7.3(11) × 10−2 [179]0.56(1)
6789.30(4)(3)484.12(7)60.0(77) × 10−3 [179]0.90(1)

a Spectroscopic factor C2 S = 0.022 from Endt and Rolfs [192], single-particle proton width calculated from the parameters given by Iliadis [193]. An uncertainty on the spectroscopic factor of 20% has been assumed. b Expectation and variance of the resonance strength using ω γ = 4.5 × 10−22 and a factor 1.5 uncertainty in [184]. c Expectation and variance of the resonance strength using ω γ = 2.1 × 10−11 and a factor 1.5 uncertainty in [184]. d This resonance is given as f0 = 0.71 in [182] citing then yet-to-be-published paper by Endt, de Wit and Alderliesten which is presumably [194]. The decay branches listed in [194] give f0 = 0.0041(7). The listed γ-ray branches in the ENSDF database give f0 = 0.0042(6). We adopt f0 = 0.0042(6). This resonance is rather weak for both radiative capture to the ground state and isomeric state and the astrophysical impact of this reassignment is negligible. e Weighted average with additional estimate of systematic uncertainty of ω γ = 9.0(6) × 10−7 [187] and ω γg = 1.1(2) × 10−7 and f0 = 0.74(1) [191]. f Weighted average 3.07(13) × 10−2 [195] and ω γ = 2.1(2) × 10−2 and f0 = 0.87(1) [191]. g f0 = 0.91(1) using ENSDF values.

The first studies on excited states above the proton threshold in 26Al were carried out via 25Mg(3He,d) reactions with the Tandem Van de Graaff accelerator at the University of Pennsylvania in late 1970s [196]. Several states were discovered and angular momenta assigned for the proton transfers based on the angular distribution of the cross sections and distorted-wave Born approximation analysis. Champagne et al did a thorough investigation of the 25Mg(3He,d) reaction at the Wright Nuclear Structure Laboratory MP Tandem Van de Graaff accelerator in early 1980s [197, 198]. States at 6343 and 6400 keV were assigned as 3+, the latter only tentatively. Proton widths and resonance strengths were deduced for many of the states by scaling from the determined proton width for the 374 keV resonance. The 25Mg(3He,d) reaction was revisited at the Princeton AVF cyclotron, and it was found that the 6343 keV state is actually the 6364 keV (3+) state and the 6400 keV is a J = 2 state at 6398 keV [199]. The state reported at 6343 keV in [197] was observed using the 27Al(3He,α)26Al but had been misassigned as a 25Mg+p resonance. It was not populated in the 25Mg(3He,d) reaction in [199], suggesting that the 6343 keV state is not a strong single-proton state and therefore does not play a crucial role in the overall reaction rate. The proton width for the 374 keV resonance was re-determined, using the γ-ray branching ratio together with previous data, in [199], and found to be significantly lower, Γp = 0.82(20) eV [199], than the previously determined value Γp = 460(129) eV [198]. The calculated resonance strength for the 92 keV resonance (ω γ ≤ 2.7 × 10−13 eV [199]) is significantly lower than the recently measured value, suggesting that the proton width might have been underestimated for the 374 keV resonance in [199]. Indeed, the resonance strength for the 92 keV resonance was revised to 8.5 × 10−11 eV in [200]. This is already much closer to the recently measured value ω γ = 2.9(6) × 10−10 [187].

The doublet of states at 6343 and 6364 keV was confirmed in [201], where Jπ = 2 was proposed for the 6399 keV state. A spin of (1+, 2) for the 6399 keV state has been adopted in the latest Nuclear Data Sheet [1] instead of the previous suggestions of 3+ [197, 198] and 2 [199, 201]. More recently, angular distribution data, measured with the Gammasphere detector array for the two most intense γ-ray transitions from the 6399 keV state, support a Jπ = 2 assignment for this state [186].

The resonance widths and strengths for the 25Mg(p, γ)26Al reaction have been studied directly using a low-energy proton beam and 25Mg target, for example in [202205]. In addition, spectroscopic factors for proton-unbound levels in 26Al and their influence on stellar reaction rates have been investigated, e.g. in [206]. For low-energy resonances, the strength is almost entirely determined by the proton width Γp : $\omega \gamma =\omega \tfrac{{{\rm{\Gamma }}}_{p}{{\rm{\Gamma }}}_{\gamma }}{{\rm{\Gamma }}}\approx \omega {{\rm{\Gamma }}}_{p}$, when Γp ≪ Γγ . The proton width depends on the spectroscopic factor, C2 S, and the single-particle width Γs.p. as Γp = C2 SΓs.p.. For a nlj proton-transfer, single-particle widths scale with the center-of-mass-energy as shown in figure 2 of [184]. Spectroscopic factors can be obtained from measurements [196, 200, 206], or from shell model calculations. For example, the strength for the 58 keV resonance has been recently extracted based on spectroscopic factors determined in the 25Mg(7Li, 6He)26Al reaction at the Q3D magnetic spectrometer of the HI-13 tandem accelerator [185].

Direct underground measurements of several resonances have been reported by the LUNA collaboration. The rate has also been recently measured underground at JUNA [207] and these data are being analysed. Limata et al [195] measured the strength for the 304 keV resonance from the emitted γ-rays and found it to be 30.7 ± 1.7 meV, in good agreement with earlier work, with the exception of the AMS work of Arazi et al [191]. Limata et al also performed an AMS study and could not reproduce the result of Arazi et al [191] but instead agreed with their own γ-ray based value. They, therefore, neglected the Arazi et al [191] value in their recommended strength of 30.8 ± 1.3 meV, which averaged their result with that of the NACRE compilation [208]. This result was then used as a reference strength for a study of the resonances at 92, 130 and 189.6 keV by Strieder et al [187]. Only an upper limit could be determined for the 130 keV resonance and it is not expected to contribute to the astrophysical rate.

Endt et al [182] studied the astrophysical aspects of the 25Mg(p, γ)26Al reaction and made thorough compilations of the excited states in 26Al [209]. Iliadis et al [184] then reinvestigated the previous literature data in 1996, and reported a new suggested rate for the reaction. In particular resonances at Er = 58 and 92 keV were found to dominate the reaction rate in the temperature region 0.02–0.15 GK [184] relevant for hydrogen burning in stars. At 0.1–1.5 GK, the resonances at 190, 304, 374, and 418 keV, start to dominate the reaction rate as shown in [191].

We have recalculated the reaction rates using the RatesMC code. The resonance energy correction results in an effective increase in the resonance energies of 1.14 keV compared to the existing literature values based on ${E}_{{\rm{res}}}={E}_{{x}}-{S}_{p}$. For resonances with only upper limits for the resonance strengths, the reduced widths are estimated by randomly sampling from a Porter–Thomas distribution using the experimental upper limit as an hard cutoff value. This is not entirely statistically accurate since the experimental upper limits should be quoted to some confidence level with an associated probability density function used to derive the upper limit at that confidence level [210] (though these data are frequently lacking in the published studies). However, the effect of this is small (see [210212] for more details). The corresponding contribution plots for the resonances are shown in figures 9 and 10. The 93- and 305 keV resonances dominate over the astrophysically relevant temperature range and both resonances have been measured directly by LUNA. The 59 keV resonance strength has thus far remained inaccessible for direct measurements. The corresponding re-evaluated reaction rates to the ground state and isomeric state of 26Al are given in tables 8 and 9. In addition to the proton-capture rate, the de-excitations from the populated states in 26Al to the ground and isomeric states of 26Al need to be taken into account. The rates are typically multiplied by the corresponding ground state branching fraction f0, see, e.g. [184]. Thermal excitation between the states, discussed in section 4, complicates the situation.

Figure 9.

Figure 9. (Left) The fractional contribution of each resonance to the 25Mg(p, γ)26Alg based on the resonance parameters given in table 1. (Right) Contour heatmap of the current reaction rate uncertainties. The thick (thin) black curves are the 68% (95%) coverage limits. The green curves are median (solid) and 68% coverages (broken) for the 25Mg(p, γ)26Alg reaction rate from the compilation of Iliadis et al [213]. The decrease in the reaction rate at low temperatures is due to the shifted resonance energy computed, taking the atomic binding into account. The decrease in the reaction rate at T ≈ 0.1 GK is due to the updated resonance strength of the 93 keV resonance from LUNA and JUNA [187, 188]. The resonance strength and ground state branching fraction for the Er = 93 keV resonance are weighted averages of [187, 188] and [188, 189], respectively.

Standard image High-resolution image
Figure 10.

Figure 10. Same as figure 9 but for 25Mg(p, γ)26Alm . The decrease in the reaction rate at low temperatures is due to the shifted resonance energy computed, taking the atomic binding into account. The difference at T ≈ 0.1 GK is due to the updated information on the Er = 93 keV resonance from [187189].

Standard image High-resolution image
Figure 11.

Figure 11. Same as figure 9 but for 26Alg (p, γ)27Si. The variations in the rates are due to the updated Γp proton partial widths derived from the 26Al(d, p)27Al experiments of [107, 214, 215].

Standard image High-resolution image

As recently evident in the study of Lotay et al, there remains some uncertainty and occassional inconsistencies in the ground state branching fractions (f0) for each resonance. For example, the f0 value of the Er = 95 keV resonance was revised from f0 = 0.52 ± 0.02(stat. ) ± 0.06 (syst.) [186] to f0 = 0.76(10) [189]. Similarly, on the basis of the γ-ray branching information stored in the ENSDF database, the f0 value for the astrophysically unimportant Er = 109 keV resonance has values ranging from f0 = 0.0041(7) to f0 = 0.71 depending on the source. New, independent measurements of these ground state branching fractions are advisable in order to validate the existing results.

3.2. The destruction reaction 26Al(p, γ)27Si

Proton capture on 26Al provides the main destruction mechanism for 26Al in several stellar sites such as classical novae, convective core hydrogen burning in massive stars, and hydrogen burning in intermediate-mass AGB stars. The uncertainties in the 26Al(p, γ)27Si reaction rate at the temperatures present in these environments can result in large variations in the 26Al abundance. For example, sensitivity studies show that this uncertainty leads to variations of up to two orders-of-magnitude in AGB calculations [89]. The situation is further complicated by the existence of the 228 keV isomeric state in 26Al, which must be treated separately from the ground state at temperatures below 0.4 GK, as discussed above [216]. Thus, the 26Alg(p, γ)27Si and 26Alm(p, γ)27Si reaction rates must be determined independently to understand the destruction of 26Al in the stellar sites described above. Due to the impact of these reaction rates on the abundance of galactic 26Al, there have been a wide variety of direct and indirect measurements aimed at determining these rates.

For 26Alg (p, γ)27Si resonances at Er > 190 keV, which have less of an influence on the destruction of 26Al, the strengths are well constrained by previous direct measurements [217, 218]. The resonance around Er = 184–190 keV is an important resonance but there is some disagreement in its resonance energy. The most recent direct measurement of 26Alg(p, γ)27Si was completed in inverse kinematics at TRIUMF using the DRAGON recoil separator with an intense 26Al ion beam (∼2.5 × 109 26Al s−1). This study determined the energy and strength of the resonance that dominates the rate in classical novae to be Er = 184 ± 1 keV and ω γ = 35 ± 7 μeV, respectively [219], resulting in an increase of 20% in 26Alg production in nova models when compared with the previous unpublished values [217]. The properties of this resonance are consistent with a p-wave assignment for this resonance Jπ = (7 − 13)/2 based on the spectroscopic factor from a 26Al(3He, d)27Si measurement by Vogelaar et al [220]. A subsequent γ-ray spectroscopy study of 27Si gave a Jπ = 11/2+ assignment to a state at Ex = 7651.9(6) keV [221, 222] based on comparison with the Ex = 7948 keV state in the mirror nucleus, 27Al, corresponding to an p = 0 resonance at Er = 188.6(4) keV. A later 26Al(d, p)27Al study clarified the spin and parity of the mirror state in 27Al as Jπ = 11/2 [214]. There is now a consensus that the resonance between Er = 184 − 190 keV has a Jπ = 11/2, p = 1 assignment. The energy of this resonance remains a matter of some disagreement and there is not, at present, any way of resolving this discrepancy. To account for this we have computed the resonance energy with an inflated uncertainty following the procedure set out in [223].

While the measurement of Ruiz et al [219] significantly reduced the uncertainty in 26Alg (p, γ)27Si for nova nucleosynthesis, lower energy resonances dominate this reaction rate at the lower temperatures found in AGB and massive stars. At these lower energies, direct measurements become unfeasible with currently available 26Al beam intensities and indirect measurements are required. Specifically, the 127 keV resonance in 27Si is thought to dominate the reaction rate in these environments as it is the only known p = 0 proton-capture resonance in this energy regime [224]. This and other resonances have been studied indirectly in a variety of measurements including transfer and γ-ray transition studies (e.g. [220, 221, 225, 226]). More recently, there have been multiple studies in inverse kinematics with unstable 26Al beams including two measurements of resonances in the mirror 27Al nucleus in inverse kinematics via the 26Al(d, p)27Al reaction [107, 214] and a measurement of states in 27Si via 26Al(d, n)27Si [227]. In these studies, spectroscopic factors of states in 27Al and 27Si were measured to determine the resonance strengths of those states. The values of the strength of the 127 keV resonance obtained via the mirror studies [107, 214] are in agreement with each other, but a factor of four higher than the previously accepted value [220]. However, there are worrying discrepancies for this resonance between the resonance information reported in [107] (C2 S = 0.0102(21) and $\omega \gamma ={2.6}_{-0.9}^{+0.7}\times {10}^{-8}$ eV) and [228] (C2 S = 0.0093(7) and ω γ = 5.7(4) × 10−8 eV), i.e. a lower spectroscopic factor is calculated to result in a larger partial width and resonance strength. Based on this, some of the discrepancies between these experimental results are likely due to the theoretical treatment of the experimental data. Revisiting the original experimental studies and treating them with consistent experimental methods may help to resolve these discrepancies and will provide better constraints on 26Al destruction by proton capture. As an example, the later study of [228] notes that the results of [214] has a technical fault in the number of nodes for the computation of the transfer amplitudes. Notwithstanding these concerns, the 127 keV resonance dominates the reaction rate at relevant temperatures for 26Al nucleosynthesis in massive and AGB stars. While the 26Al(d, n)27Si measurement only yielded an upper limit on the spectroscopic factor of this state, the results were consistent with these mirror studies [227]. Clearly, a direct measurement of this resonance strength—one which obviates some of the theoretical inconsistencies of the transfer data is a high priority for future studies once more intense 26Al beams become available.

As discussed above, at temperatures below around 0.4 GK, the isomer and ground state must be treated as separate nuclei and thus the 26Alm (p, γ)27Si reaction rate should be determined independently from proton capture on the ground state. However, previous determinations of this reaction rate were typically scaled from the 26Alg (p, γ)27Si rate, despite the large spin difference between the ground (5+) and isomeric (0+) states, as little experimental data for proton capture on the isomer was available [229]. A direct measurement of the strength of the Er = 447 keV resonance in 26Alm (p, γ)27Si is currently the only direct resonance information for this reaction [230]. A measurement of the 27Al(3He, t)27Si*(p)26Alg,m and 28Si(3He, α)27Si*(p)26Alg,m reactions was performed at the Wright Nuclear Structure Laboratory to indirectly determine the 26Alm (p, γ)27Si reaction rate based on experimental information [231]. While this study was only able to put a lower limit on the reaction rate, as proton decays from states in 27Si at energies of Er ≥ 445 keV could be measured, this study confirmed that different resonances in 27Si dominate the two rates. A similar study of the 27Al(3He, t)27Si*(p)26Al reaction using the same Enge Split-Pole spectrograph, now installed at the John D Fox Accelerator laboratory at Florida State University, was recently performed detecting proton decays down to Er ≃ 300 keV in order to further improve the reaction rate determination [232].

Alongside these charge-exchange reaction studies, a complementary investigation of the γ-decaying properties of 26Alm + p resonant states was performed at Argonne National Laboratory [222, 233, 234]. In that work [222, 233, 234], a 12C(16O, n) fusion-evaporation reaction was used to populate excited states in 27Si, located above the 26Alm + p emission energy of 7691.3(1) keV [235], and the resulting γ decays were recorded with the Gammasphere detector array [236, 237]. γ decays were observed from all resonant states with energies Er ≤ 500 keV and spin assignments were obtained from angular distribution measurements. Furthermore, by examining the mirror nucleus 27Al over a suitable energy range, it was possible to propose parity assignments for key resonances. In [222, 233, 234], it was concluded that a Jπ = 5/2+ state at Er = 146.3(3) keV is likely to dominate the astrophysical 26Alm(p, γ) reaction for low stellar temperatures, while a proposed Jπ = 3/2 state at Er = 378.3(30) keV governs the rate for temperature, T > 0.15 GK. However, it should be noted that due to the very high excitation energies of 26Alm + p resonant states, the accurate matching of mirror states is extremely challenging. In particular, it is very difficult to know how mirror shifts are affected when both the 27Si and 27Al systems become particular unbound, which is the case for a number of the proposed analog pairs.

More recently, a measurement of the 26Alm (d, p)27Al reaction in inverse kinematics was performed using an isomeric 26Al beam produced at the ATLAS facility at Argonne National Laboratory [215]. Similar to the studies of Pain et al [107] and Margerin et al [214], this measurement aimed to determine spectroscopic factors of states in 26Al(d, p)27Al reactions that are mirrors of the 26Alm + p resonances in 27Si. Using this mirror symmetry, an upper limit of the reaction rate was determined, which is dominated by resonances at Er = 146 and 378 keV over astrophysically relevant temperatures. The 26Alm (p, γ)27Si reaction rate determined here is smaller than that for the ground state by an order of magnitude or more at T9 ≤ 0.3 GK, implying that destruction via the isomer is not significant in most stellar sites. However, further study of resonances in 27Si both directly and indirectly is still warranted.

We have recomputed the 26Al(p, γ)27Si reaction rates (i.e. for the ground and metastable states) using the RatesMC code [211] and the information in tables 2 and 3. The corresponding contribution plots for the resonances are shown in figures 11 and 12. Resonances for which there is no measured proton width or resonance strength have been treated by assuming that the reduced proton widths are drawn from a Porter–Thomas distribution with mean θ2 = 0.0045 with a factor of 3 variation on that mean value. This factor variation is based on the observed scatter in the mean of the Porter–Thomas distribution determined from experimental results of proton capture and scattering reactions. For more details about this assumption including the data upon which this assumption is based, see the discussion in section 5.2.1 of [210] and the experimentally determined spread in the reduced widths reported in figure 4 and section IV.B of [212]. Experimental upper limits are used to truncate the distribution where appropriate. There are a number of different sources for the resonance information for the two reactions which have been discussed in the text above. The sources for different spectroscopic information used in the calculations of the rates are given in tables 2 and 3. Of particular note in the present evaluation is the increased uncertainty in the resonance energy of the Er = 188 keV resonance due to the disagreement between the resonances energies determined with DRAGON and γ-ray spectroscopy, and the probable need for a reconsideration of the various 26Al(d, p)27Al mirror studies to try to account for systematic differences between experiments before a full re-evaluation of the 26Alm (p, γ)27Si reaction with realistic uncertainties.

Table 2. Recommended excitation energies (Ex ) together with the spins and parities (Jπ ) for the excited states above the proton separation energy (Sp = 7463.34(13)) keV [181]) in 27Si for the 26Alg (p, γ)27Si reaction. The resonance energies (Eres) and experimentally determined resonance strengths (ω γ) for the relevant states are given where available. The atomic shift for this reaction is ΔBe = −1.27 keV. The states listed in brackets are tentative resonance states which were included in the evaluation of the reaction rate. As demonstrated in figure 11, the impact of these resonances on the reaction rate is negligible.

Ex (keV) Jπ Eres (keV) ω γ (eV)
7469.2(6)(1/2, 5/2)+ 6.8(9) < 1.8 × 10−63 [226]
(7493.1(40))(3/2+)(31(4)) < 1.5 × 10−28 a
7531.5(5)5/2+ 69.3(7) < 3.6 × 10−16 [228]
(7557(3)) [226](3/2+)(95(3)) < 3.4 × 10−15 [107]
7589.89(12)9/2+ 128.1(9)4.2(16) × 10−8 b
7651.68(11)11/2+ 187.7(23)35(7) × 10−6 [219]
7693.8(9)5/2+ 231.8(9) < 1.0 × 10−5 [217]
7704.3(2)7/2+ 242.3(2)1.0(5) × 10−5 [217]
7739.06(11)9/2+ 277.01(17)3.8(10) × 10−3 [238]
7831.5(5)9/2 369.5(5)65(18) × 10−3 [238]
8156(2) 694(2)51(27) × 10−3 [238]
8167.3(12)(11/2+)705.3(12)16(6) × 10−3 [238]
8224(2)(7/2+)762(2)35(13) × 10−3 [238]
8287(3)(7/2+ to 13/2+)825(3)41(16) × 10−3 [238]
8356(2)(3/2+ to 9/2+)894(2)67(28) × 10−3 [238]

a Resonance strength upper limit obtained using the Wigner limit for the proton width. b Weighted average of [107, 228].

Table 3. Recommended excitation energies (Ex ) together with the spins and parities (Jπ ) (taken from [234]) for the excited states above the proton separation energy (Sp = 7691.65(13)) keV [181]) in 27Si for the 26Alm (p, γ)27Si reaction. The resonance energies (Eres) and experimentally determined resonance strengths (ω γ) for the relevant states are given where available. The atomic shift for this reaction is ΔBe = −1.27 keV.

Ex (keV) Jπ Eres (keV) ω γ (eV)
7693.8(9)5/2+ 3.5(9) < 2.90 × 10−86 [239]
7704.3(2)7/2+ 14.0(2) < 4.61 × 10−44 [239]
7739.3(4)9/2+ 49.0(4) < 2.69 × 10−22 [239]
7794.8(19)7/2+ 104.5(19) < 1.92 × 10−14 [239]
7831.5(5)9/2 141.2(5) < 2.39 × 10−14 [239]
7837.6(2)5/2+ 147.2(2) < 1.5 × 10−8 [240]
7899.0(8)5/2+ 208.7(8) < 1.61 × 10−5 [239]
7909.1(7)3/2+ 218.8(7) < 1.4 × 10−6 [240]
7966.3(8)5/2+ 276.0(8) < 2.08 × 10−4 [239]
8031.5(11)5/2+ 341.2(11) < 3 × 10−8 [215]
8069.6(30)3/2 379.3(30) < 3.3 × 10−4 [240]
8139.3(6)1/2 449.3(6)434(214) × 10−3 [230]
8318(3)1/2+ 626(3)0.35(7) [231]
8375(3)1/2/3/2 683(3)0.24(6)/0.48(12) [231]
8446(3)Assumed 1/2+ 754(3)0.32(5) [231]

For the 26Alm (p, γ)27Si reaction rate, the rate has been recalculated only using known level information. While this information is somewhat lacking in the region of interest, the recent direct measurement of the 447 keV resonance strength, which dominates the reaction rate from 0.3 to 2.5 GK [230], now allows for a more accurate reaction rate determination. Nevertheless, some caution is required in the interpretation of the 26Alm (p, γ)27Si reaction rate since many of the known states in 27Si have spins and parities identified from fusion-evaporation reactions [221]; these reactions introduce a large amount of angular momentum and low-spin states may readily be missed in these experiments. Studies of low-spin states are therefore encouraged, especially indirect measurements of the 218 keV resonance, which is likely to dominate the reaction rate at temperatures below 0.3 GK. Comparison with the mirror nucleus, 27Al, which has been rather well studied may help to rule out the existence of additional states.

3.3. Neutron-induced destruction reactions

The sensitivity study by Iliadis et al [241] has shown that the 26Al(n, p)26Mg and 26Al(n, α)23Na reactions are of importance for the determination of 26Al abundances produced during hydrostatic C-shell and explosive Ne/C-shell burning phases in massive stars. These reactions involve excited states in 27Al within about 500 keV above the 26Al + n threshold (Sn = 13057.91 (12) keV) where the level density ρ is extremely high ($\rho \equiv {\rm{d}}{N}({E}_{x})/{{\rm{d}}{E}}_{x}\gt 80$ MeV−1 [242]). These states decay by proton or α-particle emission because of the lower 26Mg + p and 23Na + α thresholds (Sp = 8271 keV, Sα = 10092 keV).

At present, only few experimental data on both reactions are available. The first measurement of the 26Al(n, p1) 38 reaction was performed by Trautvetter and Käppeler using a quasi-Maxwellian neutron spectrum at around kT = 31 keV [243]. This was followed by a more comprehensive study in 1984 [244] using neutron spectra around various energies (40 meV, 31 keV, 71 keV, and 310 keV), including also the 26Al(n, p0) channel, which is 3-100 times weaker than (n, p1), and providing an upper limit on the 26Al(n, p2) channel at 40 meV. Roughly 10 years later, Koehler et al [245] determined 26Al(n, p1) and 26Al(n, α0) cross sections using the neutron time-of-flight technique at the Los Alamos Neutron Science Center (LANSCE). The 26Al(n, p1) cross section was found to be in disagreement with the earlier data [244] in the limited neutron energy range of overlap (at around 30 keV), leading to a higher stellar reaction rate by a factor of about two. Roughly 15 years ago, De Smet et al reported a measurement of 26Al(n, α0+1) reaction cross sections using the GELINA neutron time-of-flight facility at the Geel Joint Research Centre of the European Commission [246]. The GELINA measurements overlapped in the lower neutron regime with the the LANSCE experimental study (the maximum neutron energy for the (n, α0) study was 10 keV [245]) and also data at higher neutron energies were obtained. De Smet et al identified several new resonances for the 26Al(n,α) reaction. Both, Koehler et al and De Smet et al, provided resonance strengths for a resonance at around 6 keV neutron energy, however, their results disagree by a factor of 1.8. This leads to a large discrepancy of the astrophysical reaction rates deduced from these data.

Additional data on states above the neutron separation threshold and resonance strengths were obtained by Skelton et al [247], who performed an experiment using the time-reverse 26Mg(p, n)26Al and 23Na(α, n)26Al reactions, thereby accessing information on the (n, p0) and (n, α0) channels. The (n, p1) channel is thought to be astrophysically dominant, but it is not accessible in time-reversed experiments.

Recently, there have been new direct measurements of 26Al(n, p) 26Mg and 26Al(n, α) 23Na reactions at the new high neutron flux beamline EAR-2 at n_TOF CERN, and at the GELINA facility [248, 249]. Both reaction cross sections were measured up to about 150 keV neutron energy, extending the previously available experimental range for energy dependent data. Resonance strengths of several resonances were provided for the first time. Astrophysical reaction rates, including all relevant branches could be deduced from the data up to about 0.6 GK stellar temperature. Resonance strengths obtained for the 26Al(n, α) 23Na reaction in [249] agree well with previous data by De Smet et al [246], leading to a good agreement of astrophysical reaction rates at low stellar temperatures. For the 26Al(n, p) 26Mg channel astrophysical reaction rates are higher in the energy region of overlap compared to Trautvetter et al [244], however compatible within 2 standard deviations. Resonance strengths for both reactions are lower than results by Koehler et al [245].

A summary of experimental resonance strengths for 26Al + n resonances is listed in table 4 (note that these data refer to reactions on the experimentally accessible ground state of 26Al).

Table 4. Resonance energies ER and resonance strengths ω γ determined in previous direct and time-reversed experiments for 26Al(n, α) and 26Al(n, p1) resonances.

Author ER (keV) ω γα (eV) ω γα0 (eV) ω γp1 (eV)
Lederer-Woods et al [248, 249]5.9 ± 0.14.25 ± 0.414.04 ± 0.391.28 ± 0.20
 21.9 ± 0.21.62 ± 0.411.56 ± 0.40 < 0.6
 31.4 ± 0.41.62 ± 0.63 5.8 ± 1.5
 35.7 ± 0.43.7 ± 1.0 < 0.5543.4 ± 10.7
 41.3 ± 0.419.1 ± 3.69.0 ± 2.022.9 ± 5.3
 57 ± 21.8 ± 1.2 2.7 ± 1.8
 75 ± 2  8.1 ± 3.7
 86 ± 48.9 ± 7.7 85 ± 23
  ≈10538 ± 11 53 ± 14
  ≈12034 ± 10 46 ± 13
  ≈140151 ± 30 71 ± 23
     
De Smet et al [246]5.87 ± 0.024.23 ± 0.363.68 ± 0.34 
 21.98 ± 0.11.83 ± 0.271.83 ± 0.27 
 34.95 ± 0.25.98 ± 0.86only α1  
 41.3 ± 0.220.19 ± 2.0211.1 ± 1.5 
 85.2 ± 0.8   
 108.5 ± 1.1   
    
Koehler et al [245]5.578 6.6 ± 1.72.03 ± 0.51
 33.7  128 ± 22
    
     
Skelton et al [247]5.8 ± 2  ≤6.4 
 22.4 ± 2  ≤2.5 
 42 ± 2 14 ± 1.4 

Indirect studies have also been undertaken to determine the properties of 27Al states above the 26Alg + n and the 26Alm + n thresholds, e.g. excitation energy, spin/parity, branching ratio. 27Al states have been populated using proton inelastic scattering at the Tandem-ALTO (Orsay, France) and MLL (Munich, Germany) facilities. Protons were detected using the Enge Split-Pole [250] and Q3D [251] high-resolution magnetic spectrometers as the main detection system, respectively. Energies of more than 30 new 27Al states have been determined above the neutron threshold up to an excitation energy of 13.8 MeV [250]. Proton inelastic scattering has a very unselective reaction mechanism that is well adapted to populate all excited states [252254], however, on its own, it does not easily allow to identify low orbital momentum neutron capture resonances. Based on the energy of populated 27Al states, the first four lowest energy resonances observed by De Smet et al [246] were populated [250]. Proton and α-particle branching ratios have also been determined for these states by coupling a DSSSD (Double-sided Silicon Stripped Detector) array to the Enge Split-Pole spectrometer [251].

At present, the main remaining challenge is to extend our present knowledge of the properties (energy and strength) of the dominating resonances up to about 500 keV above the 26Al + n threshold. In the alternative approach of indirect measurements it may be interesting to complete the branching ratio measurement by the study of the 26Al(d, p)27Al reaction. This could be used to determine the neutron width, hence providing an indirect determination of the resonance strength. The interest of such transfer reaction is to have the same selectivity as the neutron capture reactions. Such measurements have been reported [107, 214], however, they only cover excitation energies lower than 12 MeV in 27Al. Extending these studies to higher energies would be interesting, though challenging due to the increasing background, the low proton energies, and the limited 26Al beam intensity.

3.4. The bypass reaction 25Al(p, γ)26Si

Understanding the 25Al(p, γ)26Si astrophysical reaction rate is especially important in higher-temperature environments such as novae, where this reaction becomes faster than the 25Al β decay (7.2 s). If the proton-capture rate on 25Al is faster than β decay, 26Si will be produced, which in turn decays to the 26Al isomer. The isomer subsequently decays (t1/2 = 6.3 s) to the ground state of 26Mg, and production of the long-lived 26Al ground state is bypassed (figure 6).

As radioactive beams of 25Al are not currently available at sufficient intensities to directly measure the reaction cross section, estimates of the astrophysical reaction rate have generally been based upon knowledge of the nuclear structure of 26Si. Of relevance to the rate are nuclear levels near the proton threshold at Sp = 5513.8(5) keV in 26Si [255]. Especially important is identifying 2+ or 3+ levels near the threshold that could provide s-wave resonances for the 25Al(p, γ)26Si reaction. Shell model calculations and comparisons with the 26Mg mirror nucleus indicate that levels of interest may include two 4+ states, a 1+ state, a 3+ state, and a 0+ state in the rough excitation energy range Ex = 5400–6200 keV [184, 256, 257].

Some of the first studies searching for relevant 26Si states utilized measurements of the 28Si(p, t)26Si [258] and 29Si(3He,6He)26Si reactions [259]. Energies for astrophysically-important levels were extracted from extrapolations of lower-lying level energies previously measured with high precision using γ rays [260]. The energies of several levels near threshold were determined with greater precision, and triton angular distributions from the 28Si(p, t)26Si reaction provided sensitivity to the angular momentum transfers. In 2004, Parpottas et al [261] studied the 24Mg(3He, n)26Si reaction and largely confirmed the excitation energies extracted previously [258, 259]. More interestingly, however, Parpottas et al [261] deduced, from comparisons with statistical model calculations, that the previously-observed level at ∼5916 keV was actually the important 3+ level providing an s-wave resonance in the 25Al(p, γ)26Si reaction. This hypothesis was later deemed to be consistent with the angular distribution measured in the 28Si(p, t)26Si reaction [262].

As this 3+ resonance dominates the 25Al(p, γ)26Si astrophysical reaction rate, the next most important factor, after its energy, to be determined is its resonance strength. Since the proton width is expected to be much larger than the γ width [184], the resonance strength can be determined once the γ width is known. Early estimates simply assumed the width was the same as the mirror 26Mg level (i.e. Γγ = 33 ± 14 meV) where the uncertainty comes from the uncertain 26Mg lifetime and does not account for uncertainties in isospin symmetry [184, 262]. This lifetime was recently remeasured by [263] resulting in Γγ = 33 ± 5 meV for the mirror level. Other estimates have come from combining the proton width estimated from a 25Al(d, n)26Si proton-transfer measurement [264] with subsequent determinations of the γ to proton branching ratios, Γγ p , to extract Γγ = 39 ± 21 meV [265], Γγ = 59 ± 29 meV [257], and Γγ = 71 ± 32 meV [266].

The second most important resonance contribution arises from the 1+ state at ∼5675 keV [255]. The resonance is expected to dominate at lower nova temperatures (T < 0.2 GK). The existence of this state was first identified by Caggiano et al [259], and was later verified in 24Mg(3He, n)26Si studies detecting neutrons [261] and γ rays [267269]. The strength of this resonance is determined by its proton width and estimates currently rely on shell model calculations [256] as only upper limits have been obtained for the strength of the mirror level in 25Mg(d, p)26Mg studies [270].

An additional open question has been recently highlighted by studies of the 24Mg(3He, n γ)26Si reaction [268, 269, 271]. Population of a new level at 5890 keV has been observed and conclusively identified as a 0+ state [269]. This seemingly presents an open issue since the 5949 keV level had previously been assigned as 0+ by Parpottas et al [261], and shell model calculations indicate that there should only be one 0+ state in this energy range [241, 256, 257]. Various authors have speculated that the 5949 keV may in fact be the expected 4+ level [255, 257], but this would be at odds with the cross section comparison made by Parpottas et al [261].

Figure 12.

Figure 12. Same as figure 9 but for 26Alm (p, γ)27Si, except the green lines are for 26Alg (p, γ)27Si since [213] suggests using this rate as a proxy for the 26Alm (p, γ)27Si reaction rate due to the paucity of spectroscopic information in 27Si.

Standard image High-resolution image

Recommended resonance values are displayed in table 5 and our newly calculated rate in figure 13. Considering the current state of our knowledge, it appears that the astrophysical 25Al(p, γ)26Si reaction rate is uncertain by roughly a factor of 3 at nova temperatures, primarily due to the uncertain Γγ of the Jπ = 3+ resonance. Perhaps this could be improved by directly measuring the lifetime of the 26Si 3+ level. Measurements of the neutron spectroscopic factor of the 26Mg Jπ = 1+ mirror state would also be useful to constrain the contribution of this state to the reaction rate. A repetition of the 25Mg(d, p)26Mg reaction study [270] at higher energies may be useful to improve the direct to compound nuclear component of the cross section.

Figure 13.

Figure 13. Same as figure 9 but for 25Al(p, γ)26Si. The main difference here is that the Er = 162 keV resonance is treated as an upper limit rather than as having a measured resonance strength.

Standard image High-resolution image

Table 5. Recommended excitation energies (Ex ) together with the spins and parities (Jπ ) for the excited states above the proton separation energy (Sp = 5513.98(13)) keV [181]) in 26Si. The resonance energies (Eres), proton widths (Γp ), and experimentally determined resonance strengths (ω γ) for the relevant states are given. The atomic shift for this reaction is ΔBe = −1.27 keV. Only states producing resonances below ${E}_{{\rm{res}}}=500\,\mathrm{keV}$ are listed.

Ex (keV) Jπ Eres (keV)Γp (eV) ω γ (eV)
5675.2(14) [255]1+ [269]162.5(14) <1 × 10−8 [270] <2.6 × 10−9 [270]
5890.0(8) [255]0+ [269]377.3(8)4.2(13) × 10−3 [270] a 2.4(7) × 10−4 [270] a
5928(1) [255]3+ [261]415(1)2.9(10) [264]2.3(13) × 10−2 [265]
5950(5) [255](4+) [257]437(5)7.8(39) × 10−3 [257] b 4.5(23) × 10−3 [257] b

a A 30% uncertainty was assumed for properties extrapolated from the mirror [272]. b A 50% uncertainty was assumed for properties calculated with the shell model [256].

This discussion of the reaction rate uncertainties assumes that we have a good understanding of the 26Si level structure. There are a number of open questions, however, that stretch our ability to make this claim. For instance if the 5928/5890 keV states are indeed a 3+/0+ doublet, why was the 5928 keV state populated so strongly in the 28Si(p, t)26Si reaction [258], while the 0+ state was populated so weakly? One would expect natural parity states to be populated much more strongly. Note that in [258], the 5928 keV state was labeled as 5916 keV as a result of the uncertain calibration of lower-lying levels [273], and it would be worthwhile to revisit the data from that measurement in light of the more precise energy measurements currently available. Another question is why is the 5890 keV 0+ state so readily observable in the 24Mg(3He, n γ)26Si studies [266, 268, 269, 271], but no neutrons were observed populating the state in the 24Mg(3He, n)26Si measurement [261], which seemingly had the resolution to observe it? Finally, the question as to whether we are seeing too many 0+ levels and the assertion in [263] of a missing 1 level clearly strains our ability to claim complete knowledge of the relevant 26Si level structure.

3.5. Other reactions affecting the 26Al abundance

The sensitivity study by Iliadis et al [241] identified a number of other reactions that, while not direct producing or destroying 26Al, do affect its final abundance in massive stars. The most influential are the 25Mg(α, n)28Si, the 23Na(α, p)26Mg, and the neutron capture reactions 24Mg(n, γ)25Mg and 25Mg(n, γ)26Mg. We discuss these reactions separately in the sections below.

3.5.1.  25Mg(α, n)28Si

The 25Mg(α, n)28Si reaction is most influential during explosive Ne/C-shell burning, at a temperature around 2.3 GK (Ecm = 1.8–3.5 MeV). It acts as a neutron source for the neutron-induced reactions described above, as well as reducing the 25Mg available for proton capture to 26Al.

The reaction rate currently reported in REACLIB [274] comes from the NACRE compilation [229]. Above 2 GK, the rate comes from Hauser–Feshbach calculations and below this temperature it is based on the direct measurements of Van der Zwan and Geiger [275], Anderson et al [276] and Wieland [277]. Between 0.86 and 3 MeV, only the Wieland data are used, even though this work is unpublished. However, Iliadis et al [241] questioned both the decision to use HF rates above 2 GK, as experimental data are available, and the exclusion of the Anderson et al [276] data. They further recommended a reanalysis of the available data as well as a new measurement. Several new measurements have been made at both Notre Dame and Argonne, and these data are currently under analysis.

3.5.2.  23Na(α, p)26Mg

The 23Na(α, p)26Mg reaction is thought to be influential during convective shell carbon/neon burning at temperatures around 1.4 GK [278]. It is the second most important reaction after 12C(12C, p)23Na for production of protons, which are subsequently captured by 25Mg.

Four direct measurements of this reaction cross section have recently been performed, three in inverse kinematics [239, 279, 280] and one in forward kinematics [281]. The work of Almaraz-Calderon et al [239] used a 23Na beam on a cryogenic 4He target, and detected protons between 6.8 and 13.5 deg in the laboratory. Angle-integrated cross sections were reported for the p0 and p1 channels. Tomlinson et al [279] also utilised a 23Na beam on a 4He target, this time at room temperature. Angle-integrated cross sections were given for the p0p2 channels. Avila et al [280] used the active target detector, MUSIC, to determine total cross sections. Finally, the forward kinematics measurement of Howard et al [281] extracted angular distributions for the p0 and p1 proton channels. All measurements were consistent with the Non-Smoker [282] cross sections above 1.75 MeV in the centre of mass. Finally, Hubbard et al [283] corrected the cross sections from Almaraz-Calderon et al [239] and Tomlinson et al [279] to account for the proton angular distributions measured by Howard et al [281] and calculated a new combined rate with a total uncertainty of 30% for temperatures relevant to 26Al production. This reaction rate is now considered to be sufficiently well constrained for this purpose.

3.5.3.  24Mg(n, γ)25Mg and 25Mg(n, γ)26Mg

Iliadis et al [278] identified the 24Mg(n, γ) reaction as the most important radiative neutron capture reaction impacting 26Al abundances. For example, during explosive C/Ne burning a 2 times higher 24Mg(n, γ) rate is predicted to yield a change in 26tAl abundance by a factor 1.6. Moreover, a correct interpretation of isotopic ratios in stardust spinel grains, which contain both Mg and Al (section 1.4), requires an accurate 25Mg(n, γ)26Mg reaction cross section as both this reaction and the 26Al decay lead to the production of 26Mg.

The Maxwellian-averaged cross sections (MACS) for these reactions were recommended by the Karlsruhe Astrophysical Database of Nucleosynthesis in Stars (KADoNiS) [284], quoting Maxwellian-averaged neutron capture cross sections with large uncertainties. These MACS values are based on one measurement from the 1970s [285]. Recently, high precision measurements of neutron capture on Mg isotopes have been performed at the n_TOF facility at CERN [286, 287]. In [286], resonance energies, spins and partial widths are presented for resonances in the 24Mg(n, γ) reaction up to neutron energies of about 660 keV. By using the cross section reconstructed from these resonance parameters obtained in the resolved resonance region (RRR) and the small contribution to the cross section data from evaluations at higher energy, i.e. in the unresolved resonance region (URR) [288], it was possible to determine a reliable MACS up to about kT = 300 keV, which corresponds to stellar temperatures of 3.6 GK, covering the neutron energy range of interest for 26Al synthesis. The results for different temperatures not given in [286] are listed in table 6, including the contribution of p-wave direct radiative capture [289].

Table 6. Maxwellian-averaged capture cross sections of 24Mg(n, γ) at temperatures higher than those reported in [286]. The contributions from the unresolved resonance region and the direct radiative capture are given separately. The uncertainty on the DRC contribution is mainly related to the uncertainty on the spectroscopic factors of the low-lying states populated by the direct transitions and is estimated to be approximately 20%.

TemperatureMACSReaction
kT T9 RRRURRDRCTotalRate
(keV)(K)(mb)(mb)(mb)(mb)cm3 mol−1 s−1
1201.42.3 ± 0.20.00.32.6 ± 0.37.48 × 10+05
1401.62.0 ± 0.20.00.32.3 ± 0.37.40 × 10+05
1601.91.8 ± 0.20.00.32.1 ± 0.27.29 × 10+05
1802.11.6 ± 0.20.00.32.0 ± 0.27.18 × 10+05
2002.31.4 ± 0.10.00.41.8 ± 0.27.06 × 10+05
2202.61.3 ± 0.10.10.41.7 ± 0.26.94 × 10+05
2402.81.2 ± 0.10.10.41.6 ± 0.26.82 × 10+05
2603.01.1 ± 0.10.10.41.5 ± 0.26.71 × 10+05
2803.21.0 ± 0.10.10.41.5 ± 0.26.60 × 10+05
3003.50.9 ± 0.10.10.41.4 ± 0.26.51 × 10+05

4. Decay rate of 26Al in stellar environments

While the reaction rates for the production and destruction of both 26gAl and 26mAl can be experimentally determined in terrestrial settings (section 3), they do not adequately reflect the interplay of the two states in a stellar environment. The Maxwellian temperature distribution in these scenarios leads to possible shifts in the isomeric and ground state abundance distribution via thermally excited, short-lived states (figure 14). Such a change in the final abundances would, among other effects, have direct implications for γ-ray observations as described in section 1.1. As alluded to in section 3, stellar nucleosynthesis codes can reproduce this behavior by treating the ground and isomeric states as two independent species, and modifying their respective transition, reaction and decay rates. The exact calculation of the rates shown in this work is explained in further detail in [290].

Figure 14.

Figure 14. The large number of transitions to and from short-lived, intermediate states that become available in stellar environments (left side) are being simplified by only explicitly calculating the coupling rates between all long-lived states, but including the effect the short-lived states have on them (right side). The dashed lines indicate the additionally calculated rates if the 3+ state is considered as long-lived as well, which might be relevant in certain scenarios.

Standard image High-resolution image

The distribution into the excited states, and thus the calculated transition and decay rates, are highly temperature dependent. This applies to both the rates themselves, and to the number of short-lived states that need to be considered depending on the maximum temperature of the specific astrophysical scenario. Figure 15 compares the cases of calculating the transition rates between the ground state and isomeric state by considering either just the short-lived state at 417 keV (system of 3 states), or both states at 417 and 1058 keV (see figure 14). While the transition rates agree for lower energies, a noticeable difference can be seen above temperatures of ∼50 keV. This illustrates the importance of considering the specific case for which such a rate will be used, and the upper temperature limits of achieving accurate rates with less complex systems of considered states. However, as a rule of thumb, the more complex system will never be less accurate in itself, given accurate input parameters.

Figure 15.

Figure 15. Effective coupling rates between the two long-lived states of 26Al under stellar conditions. Increasing the number of states above the isomer to be considered in the calculation as possible bypass paths from one (green) to two (blue) shows good agreement for the transition rates from ground state to isomer (top panel) and isomer to ground state (bottom panel) at lower energies. At ∼50 keV, the calculated rates start to diverge.

Standard image High-resolution image

Additionally, and probably less obviously, the categorization into long- and short-lived states is not static, either, but depends on the stellar scenario and average time-stepping of the nucleosynthesis code describing it. In the case of 26Al, there is the established isomeric 0+ state at 228 keV; but for extreme cases, the 3+ state at 417 keV with t1/2 = 1.2 ns could be considered as an isomeric state as well and be treated explicitly, changing the resulting reaction and transition rates, as illustrated in figure 16. Note that the two systems being compared contain four and eleven states here, since the case of three states in figure 15 would now not contain any short-lived states as bypass paths, but rather be just a static system. While there are branchings in the calculated rates here as well, those occur at much higher temperatures, are smaller in magnitude, and are based on an extreme case to begin with. One can therefore assume that the system with just one isomeric state and four states in total already leads to accurate rates for most scenarios.

Figure 16.

Figure 16. Effective coupling rates between the two long-lived states and the short-lived isomer (third level in figure 14) for upward (top panel) and downward (bottom panel) transitions between the long-lived states. The differences in calculated rates are much smaller and occur at higher temperatures.

Standard image High-resolution image

Both ground and isomeric states of 26Al decay almost exclusively to 26Mg via positron emission. Their individual decay rates are not affected by temperature, however, the average lifetime of the nucleus is influenced by the changes in distribution between energetic states with different decay rates. Coc et al [291] compared the effective half-life, derived from the same general approach as the one described above, with the half-life calculated from analytical off-equilibrium and equilibrium decay rates, see figure 17.

Figure 17.

Figure 17. Effective half-lifes of 26Al for different numbers of isomeric states and total states. Setting the cutoff time for which to explicitly treat a state as an isomer to 1ns adds a short-lived isomeric state (third level in figure 14) to the network.

Standard image High-resolution image

Note that these effective half-lives are only a theoretical concept and do not correlate with actual transition rates that might be usable in a stellar model. In order to generate those, one would have to first set the correct parameters for the specific stellar model, mainly the cutoff energy as determined by the maximum temperature of the star, and the time cutoff leading to the number of states to be considered long-lived, which is based on the shortest time-step of the model in question.

The difference in effective half-life in the temperature region between 10 and 30 keV, depending on the exact set of input parameters chosen, underlines the importance of considering both the temperature regime and time-stepping of the stellar scenario to produce and use the appropriate rates.

5. Summary and conclusions

We have reviewed the astrophysical importance of 26Al, as demonstrated by the many different types of astrophysical problems that sprang from the variety of 26Al observations. We have described how 26Al is produced by proton captures on the stable 25Mg in a variety of different astrophysical environments, where destruction paths via proton and neutron captures as well as by-passes can also be activated hindering its production. Finally, we have detailed the nuclear properties of 26Al and the nuclear reactions that affect its production in the different stellar sites. In table 7 we collect the relevant reactions, their sites and the temperatures.

Table 7. Summary of reactions and relevant sites. Temperatures refer to the site listed and do not necessarily represent the optimum temperature for 26Al production.

ReactionRelevant sites T (GK)
25Mg(p, γ)26AlLow- and intermediate-mass AGB stars0.06–0.1
 Hydrostatic H-core burning in massive stars0.03–0.08
 Hydrostatic C-shell burning in massive stars0.8–1.2
 Explosive Ne/C-shell burning in massive stars1.9–2.8
 Novae0.2–0.4
26Al(p, γ)27SiIntermediate-mass AGB stars0.06–0.1
 Hydrostatic H-core burning in massive stars0.03–0.08
 Novae0.2–0.4
26Al + n Low- and intermediate-mass AGB stars0.2–0.4
 Hydrostatic C-shell burning in massive stars0.8–1.2
 Explosive Ne/C-shell burning in massive stars1.9–2.8
25Al(p, γ)26SiNovae0.2–0.4
25Mg(α, n)28SiExplosive Ne/C-shell burning in massive stars1.9–2.8
23Na(α, p)26MgHydrostatic C-shell burning in massive stars0.8–1.2
24Mg(n, γ)25MgHydrostatic C-shell burning in massive stars0.8–1.2
25Mg(n, γ)26MgExplosive Ne/C-shell burning in massive stars1.9–2.8

In summary, astrophysical observations of live 26Al include its presence in the Milky Way Galaxy, observed by satellite telescopes using γ-ray spectrometers that can detect the 1.8 MeV photons produced by its decay, and potentially in terrestrial archives. The latter is, however, difficult to disentangle from the 26Al produced locally via spallation reactions. It is expected that future MeV γ-ray missions based on more advanced detector technology with unprecedented line sensitivity [2] will provide us further unique constraints from 26Al on the physics of CCSNe and how material is distributed in star-forming regions and transported in the Galaxy. Astrophysical observations of extinct 26Al come from meteoritic inclusions via measuring in the laboratory excesses in its daughter nucleus 26Mg. These samples include both stardust grains, which formed in stars and supernovae, and the first solids that formed in the early Solar System (the calcium-aluminum-rich inclusions, CAIs). The presence of 26Al in the first few Myr of the Solar System's life made an impact on the evolution of the planetesimals from which the terrestrial planets formed. However, we still do not know the origin of such 26Al, nor if it is a common radioactive nucleus in planet-forming discs in the Galaxy.

Astrophysical observations and their wide implications can only be addressed by understanding how stars and supernovae produce 26Al. Specifically, massive stars are the most relevant sites to be investigated to interpret the live galactic abundance of 26Al and its extinct abundance in the early Solar System, given that they are present in star-forming regions. These massive stars eject 26Al both via winds and by their final CCSN explosions. Stellar rotation and binarity add complexity to the modelling of 26Al production in these stars and we are still in the process of exploring using 3D modelling nucleosynthetic patterns that move beyond standard current 1D modelling. These range from proton-ingestion episodes to the merger of regions of different composition. AGB stars and novae, instead, produce roughly <10% each of the total amount of 26Al in the Galaxy, and they are of relevance predominantly for the study of the origin of extinct 26Al in stardust grains. Mixing and mass loss are the major uncertainties related to their modelling. In all the sites we discussed here, nuclear reaction uncertainties play a major role in the final estimate of the 26Al yields.

Significant progress has been made experimentally in constraining the nuclear reaction rates that determine the abundance of 26Al in the astrophysical sites discussed in this review. The 25Mg(p, γ)26Al reaction is the main production reaction in all the sites discussed. The key resonances at 93, 191, and 305 keV have been measured directly, but confirmation of the 93 keV strength would be useful. The 59 keV resonance dominates at temperatures below 0.05 GK (core H-burning in massive stars), but has remained inaccessible to direct measurements. The destruction route of the 26Al ground state via proton capture has also been studied directly at temperatures above around 0.1 GK. However, the strength of the 128 keV resonance, which dominates at lower temperatures, is still not well known and represents a high priority for future direct measurements. Proton capture on the 26Al isomeric state is much weaker than on the ground state and not considered significant in most stellar sites. For the destruction reactions through neutron capture, discrepancies between existing data at low neutron energy have been largely resolved. Additional data at higher neutron energies up to about 500 keV, relevant for 26Al destruction in massive stars are still needed. The 25Al(p, γ) by-pass reaction has been constrained through transfer studies of the relevant level information, and the remaining uncertainty is dominated by that of the radiative width of the 415 keV resonance. In relation to the main reactions that indirectly affect the abundance of 26Al, for the 25Mg(α, n)28Si reaction, new measurements have been performed and it is not yet clear whether further data will be needed. Finally, both the 23Na(α, p)26Mg and 24Mg(n, γ)25Mg reactions are now considered sufficiently well constrained across temperatures relevant in massive stars.

In conclusion, the radioisotope 26Al provides us with a wealth of information about the Galaxy, its stars and supernovae as well as the early Solar System. Many of the influential reactions are, or will shortly be, sufficiently constrained and we have summarised above the remaining experimental priorities. Further sensitivity studies, based on the latest rates, such as those newly calculated rates here and using up-to-date stellar models, are required to evaluate the impact of these remaining uncertainties. Future work will also need to investigate if more reactions than those considered here may have an impact on the production of 26Al in CCSN explosions.

Acknowledgments

This paper is based upon work from the 'ChETEC' COST Action (CA16117), supported by COST (European Cooperation in Science and Technology) and the National Science Foundation under Grant No. OISE-1927130 (IReNA). AML also acknowledges support from STFC (Science and Technology Facilities Council). PA thanks the trustees and staff of the Claude Leon Foundation for support in the form of a postdoctoral fellowship. MP acknowledges the support of NuGrid, JINA-CEE (NSF Grant PHY-1430152) and STFC (through the University of Hulls Consolidated Grant ST/R000840/1), and ongoing access to viper, the University of Hull High Performance Computing Facility. MP acknowledges the support from the 'Lendület-2014' Programme of the Hungarian Academy of Sciences (Hungary). BC acknowledges the support from the ERC Consolidator Grant (Hungary) funding scheme (Project RADIOSTAR, G.A. n. 724 560), the Hungarian Academy of Sciences via the Lendület project LP2014-17, and the National Science Foundation (NSF, USA) under Grant No. PHY-1430152 (JINA Center for the Evolution of the Elements). RR and DK are thankful for support from BMBF 05P19RFFN1 and the European Research Council under the European Unions's Seventh Framework Programme (FP/2007-2013)/ERC Grant Agreement N. 615126. AK acknowledges the funding from the European Unions Horizon 2020 research and innovation program under grant agreement No. 771036 (ERC CoG MAIDEN). DB acknowledges support from the National Science Foundation under grants PHY-2011890 (University of Notre Dame), PHY-1430152 (JINA Center for the Evolution of the Elements), and OISE-1927130 (International Research Network for Nuclear Astrophysics). CLW acknowledges support from the UK Science and Technologies Facilities Council (STFC), projects ST/P004008/1 and ST/M006085/1, and the European Research Council ERC-2015-STG Nr. 677 497. AS acknowledges support from the U.S. Department of Energy through grant DE-FG02-87ER40328 (UM). JJ acknowledges support by the Spanish MINECO Grant PID2020-117252GB-I00, by the E.U. FEDER funds, and by the AGAUR/Generalitat de Catalunya grant SGR-661/2017. Finally, we thank the ChETEC-INFRA project funded from the European Unions Horizon 2020 research and innovation programme under Grant agreement No 101008324.

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

:

Table 8. The 25Mg(p, γ)26Alg reaction rate in units of cm3 mol−1 s−1 calculated with the resonance information from table 1. The lower, median and upper rates correspond to the 32%, 50%, and 68% percentiles.

T [GK]Lower limitMedian rateUpper limit
0.0102.10 × 10−33 3.10 × 10−33 4.67 × 10−33
0.0111.29 × 10−31 1.79 × 10−31 2.55 × 10−31
0.0127.04 × 10−30 8.77 × 10−30 1.10 × 10−29
0.0133.31 × 10−28 4.06 × 10−28 4.98 × 10−28
0.0141.11 × 10−26 1.37 × 10−26 1.70 × 10−26
0.0152.49 × 10−25 3.09 × 10−25 3.85 × 10−25
0.0163.85 × 10−24 4.77 × 10−24 5.94 × 10−24
0.0183.68 × 10−22 4.55 × 10−22 5.66 × 10−22
0.0201.39 × 10−20 1.72 × 10−20 2.14 × 10−20
0.0259.18 × 10−18 1.13 × 10−17 1.40 × 10−17
0.0306.63 × 10−16 8.17 × 10−16 1.01 × 10−15
0.0401.35 × 10−13 1.65 × 10−13 2.02 × 10−13
0.0504.05 × 10−12 4.72 × 10−12 5.54 × 10−12
0.0605.58 × 10−11 6.26 × 10−11 7.03 × 10−11
0.0704.42 × 10−10 4.93 × 10−10 5.52 × 10−10
0.0802.24 × 10−09 2.51 × 10−09 2.82 × 10−09
0.0908.09 × 10−09 9.08 × 10−09 1.02 × 10−08
0.1002.29 × 10−08 2.57 × 10−08 2.90 × 10−08
0.1105.55 × 10−08 6.22 × 10−08 7.00 × 10−08
0.1201.29 × 10−07 1.45 × 10−07 1.63 × 10−07
0.1303.47 × 10−07 3.84 × 10−07 4.30 × 10−07
0.1401.17 × 10−06 1.29 × 10−06 1.43 × 10−06
0.1504.34 × 10−06 4.80 × 10−06 5.31 × 10−06
0.1601.54 × 10−05 1.71 × 10−05 1.89 × 10−05
0.1801.41 × 10−04 1.57 × 10−04 1.73 × 10−04
0.2008.55 × 10−04 9.48 × 10−04 1.05 × 10−03
0.2502.19 × 10−02 2.41 × 10−02 2.66 × 10−02
0.3001.88 × 10−01 2.07 × 10−01 2.27 × 10−01
0.3508.75 × 10−01 9.54 × 10−01 1.04 × 10+00
0.4002.77 × 10+00 3.00 × 10+00 3.26 × 10+00
0.4506.80 × 10+00 7.33 × 10+00 7.90 × 10+00
0.5001.40 × 10+01 1.50 × 10+01 1.60 × 10+01
0.6004.09 × 10+01 4.35 × 10+01 4.63 × 10+01
0.7008.75 × 10+01 9.25 × 10+01 9.78 × 10+01

Table 9. The 25Mg(p, γ)26Alm reaction rate in units of cm3 mol−1 s−1 calculated with the resonance information from table 1. The lower, median and upper rates correspond to the 32%, 50%, and 68% percentiles.

T [GK]Lower limitMedian rateUpper limit
0.0105.35 × 10−34 8.21 × 10−34 1.28 × 10−33
0.0113.17 × 10−32 4.60 × 10−32 6.86 × 10−32
0.0121.62 × 10−30 2.14 × 10−30 2.86 × 10−30
0.0137.11 × 10−29 9.42 × 10−29 1.26 × 10−28
0.0142.30 × 10−27 3.13 × 10−27 4.27 × 10−27
0.0155.08 × 10−26 7.01 × 10−26 9.65 × 10−26
0.0167.82 × 10−25 1.08 × 10−24 1.49 × 10−24
0.0187.46 × 10−23 1.03 × 10−22 1.42 × 10−22
0.0202.82 × 10−21 3.89 × 10−21 5.37 × 10−21
0.0251.86 × 10−18 2.56 × 10−18 3.53 × 10−18
0.0301.35 × 10−16 1.85 × 10−16 2.55 × 10−16
0.0402.98 × 10−14 3.94 × 10−14 5.27 × 10−14
0.0501.20 × 10−12 1.43 × 10−12 1.74 × 10−12
0.0602.10 × 10−11 2.43 × 10−11 2.83 × 10−11
0.0701.87 × 10−10 2.17 × 10−10 2.54 × 10−10
0.0809.97 × 10−10 1.17 × 10−09 1.37 × 10−09
0.0903.69 × 10−09 4.33 × 10−09 5.08 × 10−09
0.1001.05 × 10−08 1.23 × 10−08 1.45 × 10−08
0.1102.52 × 10−08 2.94 × 10−08 3.44 × 10−08
0.1205.45 × 10−08 6.34 × 10−08 7.39 × 10−08
0.1301.18 × 10−07 1.36 × 10−07 1.58 × 10−07
0.1402.91 × 10−07 3.30 × 10−07 3.80 × 10−07
0.1508.37 × 10−07 9.40 × 10−07 1.07 × 10−06
0.1602.57 × 10−06 2.89 × 10−06 3.28 × 10−06
0.1802.14 × 10−05 2.41 × 10−05 2.72 × 10−05
0.2001.28 × 10−04 1.44 × 10−04 1.63 × 10−04
0.2503.47 × 10−03 3.86 × 10−03 4.31 × 10−03
0.3003.22 × 10−02 3.54 × 10−02 3.91 × 10−02
0.3501.60 × 10−01 1.74 × 10−01 1.91 × 10−01
0.4005.34 × 10−01 5.79 × 10−01 6.31 × 10−01
0.4501.37 × 10+00 1.48 × 10+00 1.60 × 10+00
0.5002.90 × 10+00 3.13 × 10+00 3.38 × 10+00
0.6008.94 × 10+00 9.59 × 10+00 1.03 × 10+01
0.7001.98 × 10+01 2.12 × 10+01 2.27 × 10+01

Table 10. The 26Alg (p, γ)27Si reaction rate in units of cm3 mol−1 s−1 calculated with the resonance information from table 2. The lower, median and upper rates correspond to the 32%, 50%, and 68% percentiles.

T [GK]Lower limitMedian rateUpper limit
0.0102.96 × 10−37 4.35 × 10−37 6.38 × 10−37
0.0118.04 × 10−36 1.18 × 10−35 1.74 × 10−35
0.0121.51 × 10−34 2.20 × 10−34 3.22 × 10−34
0.0132.11 × 10−33 3.08 × 10−33 4.49 × 10−33
0.0142.42 × 10−32 3.58 × 10−32 5.44 × 10−32
0.0152.35 × 10−31 3.75 × 10−31 7.13 × 10−31
0.0161.97 × 10−30 3.74 × 10−30 1.19 × 10−29
0.0189.41 × 10−29 3.97 × 10−28 2.25 × 10−27
0.0203.22 × 10−27 2.61 × 10−26 1.63 × 10−25
0.0254.74 × 10−24 5.55 × 10−23 3.58 × 10−22
0.0301.18 × 10−21 9.34 × 10−21 5.80 × 10−20
0.0404.37 × 10−17 7.03 × 10−17 1.17 × 10−16
0.0504.62 × 10−14 7.04 × 10−14 1.07 × 10−13
0.0605.08 × 10−12 7.58 × 10−12 1.13 × 10−11
0.0701.48 × 10−10 2.15 × 10−10 3.14 × 10−10
0.0802.00 × 10−09 2.78 × 10−09 3.92 × 10−09
0.0901.69 × 10−08 2.26 × 10−08 3.03 × 10−08
0.1001.04 × 10−07 1.35 × 10−07 1.75 × 10−07
0.1104.98 × 10−07 6.35 × 10−07 8.11 × 10−07
0.1201.94 × 10−06 2.45 × 10−06 3.12 × 10−06
0.1306.34 × 10−06 8.00 × 10−06 1.02 × 10−05
0.1401.80 × 10−05 2.27 × 10−05 2.86 × 10−05
0.1504.58 × 10−05 5.70 × 10−05 7.15 × 10−05
0.1601.06 × 10−04 1.30 × 10−04 1.61 × 10−04
0.1804.53 × 10−04 5.43 × 10−04 6.55 × 10−04
0.2001.57 × 10−03 1.82 × 10−03 2.13 × 10−03
0.2501.87 × 10−02 2.05 × 10−02 2.27 × 10−02
0.3001.21 × 10−01 1.30 × 10−01 1.40 × 10−01
0.3505.07 × 10−01 5.44 × 10−01 5.82 × 10−01
0.4001.55 × 10+00 1.67 × 10+00 1.79 × 10+00
0.4503.76 × 10+00 4.05 × 10+00 4.36 × 10+00
0.5007.65 × 10+00 8.26 × 10+00 8.93 × 10+00
0.6002.20 × 10+01 2.39 × 10+01 2.60 × 10+01
0.7004.64 × 10+01 5.05 × 10+01 5.50 × 10+01

Table 11. The 26Alm (p, γ)27Si reaction rate in units of cm3 mol−1 s−1 calculated with the resonance information from table 3.

T [GK]Lower limitMedian rateUpper limit
0.0102.81 × 10−37 4.09 × 10−37 6.02 × 10−37
0.0117.71 × 10−36 1.13 × 10−35 1.64 × 10−35
0.0121.45 × 10−34 2.12 × 10−34 3.09 × 10−34
0.0132.01 × 10−33 2.93 × 10−33 4.25 × 10−33
0.0142.16 × 10−32 3.16 × 10−32 4.55 × 10−32
0.0151.89 × 10−31 2.74 × 10−31 3.95 × 10−31
0.0161.37 × 10−30 1.98 × 10−30 2.87 × 10−30
0.0184.51 × 10−29 6.51 × 10−29 9.37 × 10−29
0.0208.95 × 10−28 1.30 × 10−27 1.87 × 10−27
0.0253.42 × 10−25 4.99 × 10−25 7.20 × 10−25
0.0303.22 × 10−23 4.71 × 10−23 6.82 × 10−23
0.0402.69 × 10−20 3.84 × 10−20 5.50 × 10−20
0.0504.82 × 10−18 9.89 × 10−18 2.16 × 10−17
0.0603.00 × 10−16 1.35 × 10−15 4.04 × 10−15
0.0701.18 × 10−14 6.09 × 10−14 1.86 × 10−13
0.0803.16 × 10−13 1.20 × 10−12 3.36 × 10−12
0.0904.56 × 10−12 1.52 × 10−11 3.41 × 10−11
0.1003.97 × 10−11 1.30 × 10−10 2.52 × 10−10
0.1102.41 × 10−10 7.58 × 10−10 1.60 × 10−09
0.1201.11 × 10−09 3.36 × 10−09 8.34 × 10−09
0.1304.23 × 10−09 1.25 × 10−08 3.50 × 10−08
0.1401.37 × 10−08 4.08 × 10−08 1.21 × 10−07
0.1503.91 × 10−08 1.19 × 10−07 3.58 × 10−07
0.1601.04 × 10−07 3.16 × 10−07 9.29 × 10−07
0.1807.30 × 10−07 1.84 × 10−06 4.79 × 10−06
0.2005.50 × 10−06 1.03 × 10−05 2.09 × 10−05
0.2503.74 × 10−04 5.71 × 10−04 8.63 × 10−04
0.3007.70 × 10−03 1.20 × 10−02 1.88 × 10−02
0.3506.99 × 10−02 1.10 × 10−01 1.74 × 10−01
0.4003.63 × 10−01 5.71 × 10−01 9.06 × 10−01
0.4501.28 × 10+00 2.02 × 10+00 3.21 × 10+00
0.5003.47 × 10+00 5.48 × 10+00 8.70 × 10+00
0.6001.49 × 10+01 2.36 × 10+01 3.74 × 10+01
0.7004.10 × 10+01 6.47 × 10+01 1.03 × 10+02

Table 12. The 25Al(p, γ)26Si reaction rate in units of cm3 mol−1 s−1 calculated with the resonance information from table 5.

T [GK]Lower limitMedian rateUpper limit
0.0109.91 × 10−38 1.46 × 10−37 2.13 × 10−37
0.0112.72 × 10−36 3.99 × 10−36 5.87 × 10−36
0.0125.11 × 10−35 7.48 × 10−35 1.10 × 10−34
0.0137.04 × 10−34 1.03 × 10−33 1.51 × 10−33
0.0147.42 × 10−33 1.09 × 10−32 1.60 × 10−32
0.0156.38 × 10−32 9.32 × 10−32 1.37 × 10−31
0.0164.57 × 10−31 6.68 × 10−31 9.79 × 10−31
0.0181.46 × 10−29 2.13 × 10−29 3.13 × 10−29
0.0202.90 × 10−28 4.24 × 10−28 6.29 × 10−28
0.0251.16 × 10−25 1.69 × 10−25 2.48 × 10−25
0.0301.12 × 10−23 1.62 × 10−23 2.37 × 10−23
0.0408.59 × 10−21 1.25 × 10−20 1.84 × 10−20
0.0501.23 × 10−18 1.85 × 10−18 2.68 × 10−18
0.0606.79 × 10−17 1.73 × 10−16 4.42 × 10−16
0.0701.84 × 10−15 9.33 × 10−15 2.83 × 10−14
0.0803.01 × 10−14 2.10 × 10−13 6.58 × 10−13
0.0903.00 × 10−13 2.37 × 10−12 7.53 × 10−12
0.1001.99 × 10−12 1.63 × 10−11 5.22 × 10−11
0.1109.48 × 10−12 7.85 × 10−11 2.52 × 10−10
0.1203.50 × 10−11 2.87 × 10−10 9.23 × 10−10
0.1301.13 × 10−10 8.62 × 10−10 2.75 × 10−09
0.1403.78 × 10−10 2.26 × 10−09 7.02 × 10−09
0.1501.55 × 10−09 5.68 × 10−09 1.62 × 10−08
0.1606.90 × 10−09 1.56 × 10−08 3.65 × 10−08
0.1801.07 × 10−07 1.65 × 10−07 2.45 × 10−07
0.2001.07 × 10−06 1.56 × 10−06 2.33 × 10−06
0.2508.26 × 10−05 1.25 × 10−04 1.93 × 10−04
0.3001.52 × 10−03 2.32 × 10−03 3.60 × 10−03
0.3501.19 × 10−02 1.82 × 10−02 2.82 × 10−02
0.4005.41 × 10−02 8.30 × 10−02 1.29 × 10−01
0.4501.73 × 10−01 2.65 × 10−01 4.10 × 10−01
0.5004.30 × 10−01 6.59 × 10−01 1.02 × 10+00
0.6001.63 × 10+00 2.49 × 10+00 3.86 × 10+00
0.7004.08 × 10+00 6.25 × 10+00 9.66 × 10+00

Footnotes

  • 29  

    All throughout the paper the notation 26Al indicates the ground state of 26Al; when relevant the notations 26Alg and 26Alm are used, respectively, to refer to the ground and isomeric states, while 26Alt refers to the total 26Al. The focus of this paper is the stellar production and ejection of the ground state of 26Al only because the half-life of the isomer of 6 s is too short to allow this nucleus to survive and be ejected by any stellar source.

  • 30  

    The detection of the atomic lines of the radioactive elements technetium (Tc) in S-type stars by Merrill [11] represents clear evidence for currently active nucleosynthesis.

  • 31  

    Isotopic ratios of the same elements are not expected to vary significantly even in different types of minerals since both isotopes have similar chemical properties that will keep their ratio constant. This is of course not true for ratios between different elements, such as the 27Al/24Mg ratio.

  • 32  

    In addition, there is a scenario that hypothesised that 26Al was produced locally by spallation reactions induced by solar accelerated particles from the young Sun [38].

  • 33  

    Note that in the case of stardust spinel (MgAl2O4), instead, the initial abundance of 26Al needs to be disentangled from the initial abundance of 26Mg. This is complicated by the fact that in single stardust spinel grains the proportion of Mg to Al may vary from the stoichiometric value of 1:2, and such variation needs to be taken into account when attempting to derive the initial 26Al/27Al ratio [61].

  • 34  

    Since the flux corresponds to dN/dt and the decay equation is ${\rm{d}}N/{\rm{d}}t=-N/\tau $.

  • 35  

    These cores are the progenitor of the ONe white dwarfs onto which accretion results in novae producing 26Al, see section 2.3.

  • 36  

    In the bottom layers of the H-ashes 26Al is destroyed by neutron-captures, with neutrons generated by the 13C(α, n)16O reaction.

  • 37  

    These stars typically end their lives in a CCSN explosion, as opposed to the low- and intermediate-mass stars discussed in the previous section that become AGB stars at the end of their evolution and shed material via stellar winds.

  • 38  

    The subscript 1 refers to particle emission to the first excited state of 26Mg.

Please wait… references are loading.