Brought to you by:

A publishing partnership

POSSIBLE RESONANCES IN THE 12C + 12C FUSION RATE AND SUPERBURST IGNITION

, , and

Published 2009 August 14 © 2009. The American Astronomical Society. All rights reserved.
, , Citation Randall L. Cooper et al 2009 ApJ 702 660 DOI 10.1088/0004-637X/702/1/660

0004-637X/702/1/660

ABSTRACT

Observationally inferred superburst ignition depths are shallower than models predict. We address this discrepancy by reexamining the superburst trigger mechanism. We first explore the hypothesis of Kuulkers et al. that exothermic electron captures trigger superbursts. We find that all electron capture reactions are thermally stable in accreting neutron star oceans and thus are not a viable trigger mechanism. Fusion reactions other than 12C + 12C are infeasible as well since the possible reactants either deplete at much shallower depths or have prohibitively large Coulomb barriers. Thus, we confirm the proposal of Cumming & Bildsten and Strohmayer & Brown that 12C + 12C triggers superbursts. We then examine the 12C + 12C fusion rate. The reaction cross section is experimentally unknown at astrophysically relevant energies, but resonances exist in the 12C + 12C system throughout the entire measured energy range. Thus it is likely, and in fact has been predicted, that a resonance exists near the Gamow peak energy Epk ≈ 1.5 MeV. For such a hypothetical 1.5 MeV resonance, we derive both a fiducial value and upper limit to the resonance strength (ωγ)R and find that such a resonance could decrease the theoretically predicted superburst ignition depth by up to a factor of 4; in this case, observationally inferred superburst ignition depths would accord with model predictions for a range of plausible neutron star parameters. Said differently, such a resonance would decrease the temperature required for unstable 12C ignition at a column depth 1012 g cm−2 from 6 × 108 K to 5 × 108 K. A resonance at 1.5 MeV would not strongly affect the ignition density of Type Ia supernovae, but it would lower the temperature at which 12C ignites in massive post-main-sequence stars. Determining the existence of a strong resonance in the Gamow window requires measurements of the 12C + 12C cross section down to a center-of-mass energy near 1.5 MeV, which is within reach of the proposed DUSEL facility.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Superbursts are long, energetic, and rare thermonuclear flashes on accreting neutron stars in low-mass X-ray binaries. Their durations (∼hours), fluences (∼1042 erg), and recurrence times (∼years) distinguish superbursts from their typical hydrogen- and helium-triggered counterparts (for reviews, see Kuulkers 2004; Cumming 2005; Strohmayer & Bildsten 2006). As of this writing, astronomers have detected 15 superbursts from 10 sources (Kuulkers 2004; in't Zand et al. 2004; Remillard et al. 2005; Kuulkers 2005; Keek et al. 2008, and references therein).

The proposal (Cumming & Bildsten 2001; Strohmayer & Brown 2002) that thermally unstable 12C fusion (Woosley & Taam 1976; Taam & Picklum 1978; Brown & Bildsten 1998) triggers superbursts offers a reasonable explanation of their origin. Cooling model fits to superburst light curves (Cumming & Macbeth 2004; Cumming et al. 2006) as well as observed fluences and recurrence times (e.g., Keek et al. 2006) suggest ignition column depths Σign ≈ 1012 g cm−2, where Σ ≡ ∫ρ dz is the radially integrated density. Previous superburst ignition models (Cumming & Bildsten 2001; Strohmayer & Brown 2002; Cumming 2003; Brown 2004; Cooper & Narayan 2005; Cumming et al. 2006; Gupta et al. 2007) demonstrated that 12C ignites at Σ ≈ 1012 g cm−2 only if 12C is abundant and the ocean temperature T ≈ 6 × 108 K at that column depth; within existing models of nuclear heating in the neutron star crust, such a large temperature requires an inefficient neutrino emission mechanism in the neutron star core and a low thermal conductivity in the neutron star crust, so that the crust is much hotter than the core.

Recent observations, simulations, and experiments have exposed three fundamental problems with this scenario. First and foremost is the inference that the ocean is in fact too cold for 12C ignition at the inferred column depth Σign ≈ 1012 g cm−2. This comes from fits (Shternin et al. 2007; Brown & Cumming 2009) to the quiescent cooling of the quasi-persistent transient KS 1731 − 260 (Wijnands et al. 2002; Rutledge et al. 2002; Cackett et al. 2006), a system that also exhibited a superburst (Kuulkers et al. 2002b). The timescale for the quiescent luminosity to decrease suggests that the crust's thermal conductivity is high; as a result, the inner crust temperature remains close to that of the core even during the accretion outburst. Cackett et al. (2008) reach the same conclusion for MXB 1659 − 29 (see also Brown & Cumming 2009). In fact, molecular dynamics simulation results (Horowitz et al. 2007, 2009; Horowitz & Berry 2009) suggest that the neutron star crust is arranged in a regular lattice and therefore has a high thermal conductivity. Neither shear-induced viscous heating (Piro & Bildsten 2007; Keek et al. 2009) nor deep crustal heating due to electron captures, neutron emissions, and pycnonuclear reactions (e.g., Haensel & Zdunik 2008; Horowitz et al. 2008; Gupta et al. 2008) can account for the heat necessary to raise the ocean temperature to the required level (although see Page & Cumming 2005; Blaschke et al. 2008, who consider heating in strange stars and hybrid stars, respectively).

Second, evidence of heavy-ion fusion hindrance at extreme sub-Coulomb-barrier energies (Jiang et al. 2002, 2007, 2008) implies that the cross section and thereby the 12C + 12C reaction rate may be orders of magnitude smaller than that assumed in the aforementioned superburst ignition models. When included in superburst ignition models, heavy-ion fusion hindrance increases Σign by at least a factor of 2 (Gasques et al. 2007).

Third, the means by which nuclear burning on the stellar surface produces sufficient quantities of 12C to trigger superbursts is poorly understood. Superburst models require large 12C mass fractions for ignition (Cumming & Bildsten 2001; Cumming 2003; Cooper & Narayan 2005; Cooper et al. 2006; Cumming et al. 2006). All systems that exhibit superbursts show helium-triggered type I X-ray bursts as well (e.g., Galloway et al. 2008), but theoretical models of such bursts yield 12C mass fractions far smaller than those required for ignition (Joss 1978; Schatz et al. 2001, 2003b; Koike et al. 2004; Woosley et al. 2004a; Fisker et al. 2005, 2008; Peng et al. 2007; Parikh et al. 2008). Most systems that exhibit superbursts apparently undergo long periods of stable nuclear burning between successive helium-triggered bursts (Kuulkers et al. 2002a; in't Zand et al. 2003; Keek et al. 2008); stable burning generates much more 12C than unstable burning, but the calculated yield is insufficient to trigger superbursts in all systems, particularly those accreting at a high rate (Taam & Picklum 1978; Schatz et al. 1999, 2003b; Cooper et al. 2006; Fisker et al. 2006).

Detection of a superburst from the classical transient 4U 1608 − 522 (Remillard et al. 2005; Kuulkers 2005) with Σign ≈ 1012 g cm−2 (Keek et al. 2008) exacerbates all three problems: (1) the transient's inferred ocean temperature is lower than those of other systems exhibiting superbursts; (2) heavy-ion fusion hindrance is greater at lower temperatures; and (3) most of the matter accreted onto the neutron star prior to the observed superburst likely burned during helium-triggered type I X-ray bursts, which current theoretical calculations suggest would generate far less 12C than that required for ignition.

Reconciling superburst observations with the current theoretical model is impossible. This motivates both a critical assessment of the current ignition model and a search for alternative ignition mechanisms.

Kuulkers et al. (2002b) proposed such an alternative mechanism. Electron captures onto protons and the subsequent captures of the resulting neutrons onto heavy nuclei liberate ≈ 7 MeV/mu (Bildsten & Cumming 1998). Pre-threshold captures of super-Fermi electrons are very temperature sensitive and therefore could trigger an energetic thermonuclear flash. An attractive feature of this mechanism is that ignition occurs always at the same electron chemical potential; thus Σign would be similar for all superbursts, in accord with observations. Unfortunately, the calculated Σign ≈ 2 × 1010 g cm−2 is much smaller than the inferred superburst Σign, making it an unlikely trigger mechanism. This motivates our investigation of exothermic electron captures onto heavy nuclei, which occur throughout the ocean and crust of an accreting neutron star, including the superburst ignition region (Sato 1979; Blaes et al. 1990; Haensel & Zdunik 1990, 2003, 2008; Gupta et al. 2007). In Section 2, we determine the thermal stability of electron captures onto heavy nuclei in accreting neutron stars. We show that instability requires unrealistically large reaction $\mathcal {Q}$-values, where $\mathcal {Q}$ is the energy released per capture; thus we conclude that electron captures in accreting neutron star oceans are thermally stable. We then consider the relevance of α captures onto light elements such as 12C in Section 3. We find that none of these reactions is a feasible mechanism and thereby confirm the proposal that 12C + 12C triggers superbursts.

In Section 4, we assess whether the 12C + 12C reaction rate could be much larger than the fiducial rate. We investigate (Section 4.1) the screening enhancement factor, including a careful evaluation of corrections to the liner mixing rule, and show that uncertainties in the plasma screening enhancement are unlikely to change the 12C + 12C reaction rate enough. We then consider the nuclear cross section. We find that a strong resonance at an energy near 1.5 MeV in the 12C + 12C system, which theoretical nuclear physics models predict, could increase the reaction rate in the astrophysically relevant temperature range by over 2 orders of magnitude. In Section 5, we show that the existence of such a resonance could decrease the predicted Σign by a factor ≈2–4 and thereby alleviate the discrepancy between superburst models and observations. We conclude in Section 6 by discussing the implications of our findings.

2. THERMAL STABILITY OF ELECTRON CAPTURES

Consider the accretion-driven compression of a matter element containing a nucleus of mass M(A, Z), where A is the mass number and Z is the proton number. The degenerate electrons' chemical potential μe rises as the nucleus advects to higher pressures. Eventually M(A, Z) + μe/c2 exceeds M(A, Z − 1) and electron capture becomes energetically favorable. Such captures often occur in equilibrium and release a negligible amount of energy; however, some captures can heat the ocean in two mutually inclusive ways (e.g., Gupta et al. 2007). (1) An electron captures into an exited state of the daughter nucleus if, for example, the daughter nucleus's ground state is forbidden. The daughter nucleus then radiatively de-excites and thereby heats the ocean. (2) If the parent nucleus is even–even, then M(A, Z − 1)>M(A, Z − 2) due to the nuclear pairing energy, and a second electron capture immediately ensues. The latter, post-threshold electron capture occurs out of equilibrium and thus releases heat.

2.1. Governing Equations

We construct a simple model of the accreted layer to determine the stability of exothermic electron captures to thermal perturbations. We assume spherical accretion onto a neutron star of mass M = 1.4 M and radius R = 10 km at an accretion rate per unit area $\dot{\Sigma }$. The accreted layer's scale height is much less than R, so we set the gravitational acceleration g = GM/R2(1 − 2GM/Rc2)−1/2 = 2.43 × 1014 cm s−2 throughout the layer. The layer is always in hydrostatic equilibrium, so the column depth Σ is a good Eulerian coordinate. To facilitate comparisons between microphysical and observationally inferred quantities, we express microphysical quantities in terms of the macroscopic coordinate Σ using the following approximate relation between mass density ρ and Σ for relativistic, degenerate electrons,3

Equation (1)

where 〈A/Z〉 is the mean molecular weight per electron and Σ = Σ12 × 1012 g cm−2. We denote the Eulerian time and spatial derivatives as ∂/∂t and ∂/∂Σ, respectively, and the Lagrangian derivative following a matter element as D/Dt, where $D/Dt = \partial /\partial t \,\,{+}\,\, \dot{\Sigma }\partial /\partial \Sigma$. The governing transport, entropy, and continuity equations are

Equation (2)

Equation (3)

Equation (4)

where F is the flux, K is the thermal conductivity, s is the entropy, $\mathcal {E}= \mathcal {Q}/(Am_\mathrm{u})$ is the energy per gram released via electron captures, X is the parent nucleus mass fraction,

Equation (5)

is the electron capture rate (Fuller et al. 1985), 〈ft〉 is the effective ft value (Fuller et al. 1980, 1985; Langanke & Martínez-Pinedo 2001), me is the electron mass, Q is the threshold energy, and

Equation (6)

is the Fermi–Dirac distribution function.

Consider pre-threshold electron captures, where μe < Q. For T = 0, all electrons have energies E ⩽ μe by Equation (6); electron capture is blocked. For T > 0, some electrons have E > Q and thus can capture. The number of electrons with E > Q increases with T, which makes pre-threshold electron capture temperature sensitive. In the pre-threshold limit (μeQ)/kBT ≪ 0,

Equation (7)

(Fuller et al. 1985; Bildsten & Cumming 1998). Conversely, for μe > Q a majority of electrons has E > Q and hence can capture for any T, making rec relatively temperature insensitive and thus thermally stable. We therefore consider pre-threshold electron captures exclusively hereafter.

2.2. Pre-threshold Electron Capture

Pre-threshold electron captures occur within a thin layer in the deep ocean. To illustrate this, consider the height-integrated capture rate. Relativistic, degenerate electrons supply the pressure P = gΣ, so

Equation (8)

Integrating Equation (7) over Σ and using Equation (8),

Equation (9)

for μekBT. Equation (9) shows that pre-threshold electron captures occur in a narrow column depth range

Equation (10)

where T = T8 × 108 K.

Now consider electron captures in steady state, such that electrons capture onto nuclei at the same rate as accretion advects the nuclei (see also the discussion in Bildsten & Cumming 1998). Equation (4) becomes

Equation (11)

Integrating Equation (11) from 0 to Q and using Equations (8) and (9), we find that most electron captures occur pre-threshold when

Equation (12)

where $\dot{\Sigma }_{\mathrm{Edd}}\approx 10^{5} \ \mathrm{g}\ {\mathrm{c}\mathrm{m}}^{-2}\ {\mathrm{s}}^{-1}$ is the local accretion rate at which the accretion flux equals the Eddington flux.4 Superbursts ignite at column depths Σ12 ∼ 1 and accretion rates $\dot{\Sigma }\approx 0.1$$1 \;\dot{\Sigma }_{\mathrm{Edd}}$. Equation (12) shows that superallowed electron captures (for which 〈ft〉 ∼ 103–104 s) occur pre-threshold at superburst ignition depths.

2.3. Thermal Stability Analysis

We now derive a one-zone model (e.g., Fujimoto et al. 1981; Paczyński 1983; Bildsten 1998b) from the governing Equations (2)–(4) to determine the stability of pre-threshold electron captures to thermal perturbations and thereby ascertain whether electron captures trigger superbursts. We consider only temperature perturbations and ignore the accretion-induced entropy advection through the bottom of the zone. Therefore, we set ∂/∂t = 0 in Equation (4) and approximate Ds/Dt = ∂s/∂t in Equation (3). Perturbations occur at constant pressure since the scale height ∼Σ/ρ ≪ R; therefore, we write Tds = CPdT, where CP is the specific heat at constant pressure. Equations (2)–(4) become

Equation (13)

Equation (14)

Equation (15)

We simplify Equations (13)–(15) as follows. We set ρ, F, and X to be constant throughout the layer; specifically, we adopt step-like profiles for F and X,

Equation (16)

where Θ is the Heaviside step function, Σec denotes the column depth at the bottom of the layer, and F0 and X0 denote the values at the top of the layer, where Σ ≪ Σec. We assume the ocean consists of a single ion species and set X0 = 1. Electron-ion scattering sets the ocean's thermal conductivity (Yakovlev & Urpin 1980; Itoh et al. 1983; Potekhin et al. 1999)

Equation (17)

where we set the Coulomb logarithm Λei = 1, a value appropriate for a plasma at T8 ≈ 5 and ρ ≈ 6 × 108 g cm−3. Since KT, we rewrite Equation (13) as

Equation (18)

Equations (16) and (18) then imply

Equation (19)

Integrating Equations (14)–(15) over Σ and using Equations (9), (10), (16), (18), and (19), we find

Equation (20)

Equation (21)

Equation (21) shows that electron capture occurs when the lifetime of a parent nucleus 1/rec equals the time taccΔΣ/Σ, an advecting element spends within the capture region, where

Equation (22)

is the accretion timescale. Note that this differs from the usual assumption (Blaes et al. 1990; Bildsten 1998a; Bildsten & Cumming 1998; Ushomirsky et al. 2000) that electron capture occurs when 1/rec = tacc.

Finally, conducting a linear stability analysis on Equation (20) and using Equation (21), the thermal instability criterion is

Equation (23)

where the temperature sensitivity of the height-integrated electron capture rate

Equation (24)

from Equations (7) and (9). Noting that electrons capture when μe/Q ≈ 1, we write Equation (24) as

Equation (25)

Using Equations (22) and (17), the thermal instability criterion (23) becomes

Equation (26)

where $\mathcal {Q}$ is the energy released per electron capture.

We tested the accuracy of Equation (23) using the suitably modified global linear stability analysis of Cooper & Narayan (2005). The minimum $\mathcal {Q}$ for instability derived from the global stability analysis differed from that of Equation (26) by less than 30% for each of the 12 test cases.

Typically $\mathcal {Q}< Q$ because the daughter nucleus is generally more massive than the parent nucleus, although exceptions exist. Gupta et al. (2007) find $\mathcal {Q}< 6.2\;\mathrm{M}\mathrm{eV}$ for all electron captures that occur for μe < 6 MeV, or equivalently, Σ12 < 10 (Equation (8)). From Equation (26), it follows that electron captures are thermally stable for the accretion rates and column depths at which superbursts occur. Therefore, we conclude that electron captures do not trigger superbursts.

3. ALTERNATIVE FUSION REACTIONS

In this section, we examine whether light-element fusion reactions trigger superbursts. Hydrogen has an electron capture threshold energy Q = 1.2933 MeV and thus depletes at Σ12 ≲ 2 × 10−2 (Equation (8)). The helium abundance at Σign is less certain, and we discuss it below. The paucity of stable isotopes of Z = 3–5 nuclei leaves 12C as the next reasonable alternative, which we address in Section 4. Finally, nuclei with Z > 6 are unlikely candidates because the extra Coulomb repulsion causes the fusion rates to be significantly lower than that of 12C.

Thus, other than 12C + 12C, α capture reactions such as 12C(α, γ)16O are the only plausible fusion reactions that might trigger superbursts. The conditions for these reactions to produce superbursts are similar to those for electron captures, namely, that (1) the reaction rate rnuc is sufficiently temperature dependent to produce unstable burning, and (2) α particles must survive to the inferred Σign. Below, we show that the latter condition is not met; thus α particles deplete too quickly to trigger superbursts.

The condition for α particles to survive at a column depth Σ is $Y / \sum _i \left(r_{\mathrm{nuc}}\right)_i > \Sigma /\dot{\Sigma }$ (see Section 2.3), where Y is the helium mass fraction and the sum is over all reactions that consume α particles. Using the triple-α reaction rate of Fushiki & Lamb (1987) and setting ρ = 6 × 108 g cm−3, T8 = 5, and Y = 1, we find rnuc = 2.2 × 106s−1, which implies a lifetime of 4.6 × 10−7s. The reaction rate rnucY3, so the lifetime is much larger for smaller helium abundances. The accretion timescale $t_{\mathrm{acc}}= 10^{7} [\Sigma _{12}/(\dot{\Sigma }/\dot{\Sigma }_{\mathrm{Edd}}) ] \;\mathrm{s}$ (Equation (22)), indicating that Y < 10−7 for helium to survive. At this low Y, the rise in temperature from consuming the helium via, e.g., 12C(α, γ), is ≲106 K ≪ T and hence insufficient to trigger a thermal instability.

From the results of this section and Section 2, we conclude that 12C fusion triggers superbursts.

4. THE 12C + 12C REACTION RATE

A possible solution to the superburst ignition problem is that the true 12C + 12C fusion rate is larger than assumed. 12C ignition at the inferred Σign requires a ∼104 reaction rate enhancement for an ocean temperature of 4 × 108 K (Cumming et al. 2006) or a ∼102 enhancement for 5 × 108 K. The two sources of uncertainty in the fusion rate are (1) plasma screening effects and (2) the nuclear cross section σ(E). In the following subsections, we investigate whether either source could account for such a large increase in the fusion rate.

4.1. Plasma Screening

Superbursts ignite in a strongly coupled Coulomb plasma. Two dimensionless parameters determine the plasma's state. The first is the Coulomb coupling parameter

Equation (27)

where a = (3Z/4πne)1/3 is the ion-sphere radius, ne is the electron number density, and we used Equation (1), which assumes the gravitational acceleration g = 2.43 × 1014 cm s−2. For Γ ≪ 1, Coulomb coupling is weak and the ions constitute a Maxwell–Boltzmann gas. As Γ increases, the ions gradually become a Coulomb liquid. When Γ>175, the ions crystallize (Potekhin & Chabrier 2000). Equation (27) implies that superbursts ignite in a Coulomb liquid. The second dimensionless parameter,

Equation (28)

is the ratio of the classical turning point to the ion separation, where

Equation (29)

and μ is the reduced mass of the reacting nuclei. Specifically, ζ = rTP/a, where rTP is the radius at which the Coulomb energy Z2e2/rTP equals the classical Gamow peak energy (Clayton 1983)

Equation (30)

Many-body interactions in a strongly coupled Coulomb plasma modify the Coulomb potential between two reacting nuclei (for a review, see Ichimaru 1993). From these many-body interactions, one derives an effective two-body potential

Equation (31)

where r is the distance between the reacting nuclei. From H(r), one derives the plasma screening enhancement to the reaction rate exp(〈H(r)〉/kBT), where 〈H(r)〉 is a path-integral average of H(r) (e.g., Ichimaru 1993). One can expand the static mean-field potential H(r) as a power series in (r/a)2 (Widom 1963). Neglecting quantum effects in H(r), the leading order term H(0) is a thermodynamic quantity; H(0) equals the difference between the Coulomb (or excess) Helmholtz free energy before and after the reaction (DeWitt et al. 1973).

Monte Carlo simulations and hypernetted chain calculations of binary ionic mixtures (Hansen & Vieillefosse 1976; Hansen et al. 1977; Chabrier & Ashcroft 1990; Ogata et al. 1993; Rosenfeld 1995, 1996; DeWitt et al. 1996; DeWitt & Slattery 2003) suggest that the excess free energy obeys the linear mixing rule to high accuracy in the regime Γ>1 (Potekhin et al. 2009). Therefore, authors usually invoke the linear mixing rule when deriving the plasma screening enhancement to the reaction rate. In this case, the total free energy of an ionic mixture

Equation (32)

where Ni is the number of ions with charge Zi, fexFex,OCP/NkBT is the well-determined reduced excess free energy per ion of a one-component plasma (e.g., Chabrier & Potekhin 1998; Potekhin & Chabrier 2000), and ΓiZ5/3i is the Coulomb coupling parameter for species i (Equation (27)).

From Equation (32), H(0)/kBT = 2fex(Γ) − fex(25/3Γ) (Jancovici 1977, see also the Appendix), where Γ is that of the reacting ions; using the ion-sphere model result fex ≈ −0.9Γ (Salpeter 1954), H(0)/kBT = 0.9(25/3 − 2)Γ ≈ 1.0573Γ, so the lowest-order screening enhancement to the 12C + 12C reaction rate

Equation (33)

Despite its simplicity, Equation (33) is adequate for most applications (Gasques et al. 2005; Yakovlev et al. 2006; Chugunov et al. 2007). We discuss corrections to the screening enhancement below.

4.1.1. Deviation from Linear Mixing Rule

The excess free energy Fex of a multicomponent plasma exhibits small deviations from the linear mixing rule (Equation (32)). In general,

Equation (34)

where Δfex ⩾ 0 is a function of both the charges Zi and concentrations xiNi/N of the ionic species (e.g., DeWitt et al. 1996). Using the hypernetted chain calculations of DeWitt et al. (1996) and the ansatz Δfexx1x2(Z2/Z1)3/2 (DeWitt & Slattery 2003), we find

Equation (35)

for a binary ionic mixture (see also Potekhin et al. 2009).

To incorporate linear mixing rule deviations into H(0) calculations, previous authors assumed a one-component plasma (consisting of 12C ions in this case). Fusion of two 12C ions generates a compound nucleus 24Mg and thereby forms a binary ionic mixture. One determines H(0) by finding the difference in Fex before and after the reaction in the limit that the compound nucleus concentration x2 → 0 (Ichimaru 1993). Using this assumption and Equations (34) and (35), we find the correction to H(0),

Equation (36)

in good agreement with DeWitt et al. (1996); for a one-component plasma, linear mixing rule deviations reduce the plasma screening enhancement factor by ∼10%.

However, the plasma at the superburst ignition depth is likely a mixture of 12C and heavier ions with Z ≲ 46 (e.g., Koike et al. 1999, 2004; Schatz et al. 2001, 2003b; Woosley et al. 2004a). Generalizing Equation (35) for a multicomponent plasma with Zi < Zj for i < j, we find (Ogata et al. 1993)

Equation (37)

To illustrate the effect spectator ions have on the screening enhancement, consider for simplicity a ternary ionic mixture of 12C, 24Mg, and a representative spectator ion 56Fe. Using Equations (34) and (37) and again taking the 24Mg concentration x2 → 0, so that x1 + x3 = 1, we find (see Appendix)

Equation (38)

Note that Equation (38) reduces to (36) in the limit x3 → 0, as it should. Equation (38) shows that, since the bracketed term is positive, heavy spectator ions increase the screening enhancement factor (i.e., ∂ΔH(0)/∂x3 > 0). For the fiducial 12C mass fraction 0.2 and 56Fe mass fraction 0.8 (such that x1 = 7/13 and x3 = 6/13; Cumming et al. 2006), linear mixing rule deviations increase the plasma screening enhancement by ≈10%.

Potekhin et al. (2009) developed an analytic formula for Δfex that is more accurate than Equation (37). We derive the corresponding formula for ΔH(0)/kBT in the Appendix. In Figure 1, we plot ΔH(0)/kBT as a function of the spectator ion number fraction x3 using both Equation (38) and the expression derived from Potekhin et al. (2009). Figure 1 confirms that heavy spectator ions increase the plasma screening enhancement to the reaction rate, although the two expressions for ΔH(0)/kBT differ quantitatively.

Figure 1.

Figure 1. Correction to the plasma screening enhancement factor due to linear mixing rule deviations as a function of the spectator ion number fraction x3. Considered is a ternary ionic mixture with Z1 = 6, Z2 = 12, Z3 = 26, and Γ1 = 6. The number fraction of the product ion x2 = 0, so x1 = 1 − x3. "PCR09" refers to the expression for ΔH(0)/kBT derived from the results of Potekhin et al. (2009; see Appendix).

Standard image High-resolution image

4.1.2. Corrections to 〈H(r)〉

The next term in the expansion of H(r) goes as (r/a)2 ∝ ζ2; its contribution is small because ζ2 ≪ 1 (Equation (28)). From Jancovici (1977), we find

Equation (39)

This result agrees very well with more accurate calculations (Alastuey & Jancovici 1978; Ogata et al. 1991; Ogata 1997). Higher order terms are even smaller; therefore, we conclude that corrections to 〈H(r)〉 are unimportant in calculating the plasma screening enhancement for 12C ignition.

4.1.3. Electron Screening Corrections

In the above analysis, we tacitly assumed a uniform electron density. Although highly degenerate, electrons nonetheless slightly concentrate around positively charged ions. Electron polarization mitigates the Coulomb repulsion between ions relative to an unpolarized configuration. This has two counteracting effects: it (1) lowers the Coulomb repulsion between the two reacting ions, which increases the reaction rate, and (2) attenuates the many-body Coulomb interactions and thereby H(r), which decreases the reaction rate. The Yukawa potential Z2e2/rexp(−r/rTF) describes the two-body potential, where rTF is the Thomas–Fermi screening length. For relativistic, degenerate electrons, rTF/a = 3.0(Z/6)−1/3 (e.g., Haensel et al. 2007), so electron screening is weak; from the results of Sahrling & Chabrier (1998), electron screening changes the reaction rate by ≲1%.

Corrections to the lowest-order plasma screening enhancement (Equation (33)) change the 12C + 12C reaction rate by a factor < 2. Therefore, we conclude that uncertainties in the plasma screening enhancement are too small to explain the discrepancy between superburst observations and theoretical model results.

4.2. The Nuclear Cross Section

Although the plasma screening enhancement to the 12C + 12C reaction rate is well determined for superburst conditions, the nuclear cross section σ(E) is not. Many groups have measured σ(E) at various center-of-mass energies E down to ≈2.1 MeV (Patterson et al. 1969; Mazarakis & Stephens 1972, 1973; Spinka & Winkler 1974; High & Čujec 1977; Kettner et al. 1977, 1980; Erb et al. 1980; Treu et al. 1980; Becker et al. 1981; Dasmahapatra et al. 1982; Satkowiak et al. 1982; Rosales et al. 2003; Barrón-Palos et al. 2004, 2006; Aguilera et al. 2006; Spillane et al. 2007). However, the energy range of interest is centered at the classical Gamow peak energy (cf. Equation (29)–(30)),

Equation (40)

and has a full width

Equation (41)

Thus, σ(E) in the astrophysically relevant energy range is experimentally unknown.

This situation is common in nuclear astrophysics: to determine the astrophysical reaction rate, one either extrapolates the experimental data to lower energies or calculates the rate theoretically (e.g., Caughlan & Fowler 1988; Gasques et al. 2005). In doing so, one tacitly assumes that the astrophysical rate has the nonresonant form, i.e., no prominent resonances exist in the compound nucleus within the relevant energy range.5 However, several groups have detected strong resonances in the 12C + 12C system at energies below the Coulomb barrier. Resonances exist throughout the entire energy range probed so far, and the spacing between adjacent resonances6 is ≈0.3 MeV. Therefore, a resonance probably exists near Epk (Bromley et al. 1960; Almqvist et al. 1960; Galster et al. 1977; Korotky et al. 1979; Erb et al. 1980; Treu et al. 1980; Spillane et al. 2007). Indeed, Michaud & Vogt (1972) and Perez-Torres et al. (2006) predict that a resonance exists in the 12C + 12C system with energy ER ≈ 1.5 MeV. If the resonance is strong, the thermally averaged reaction rate 〈σv〉 would be much larger than assumed.

To illustrate the effect a strong resonance within the Gamow window would have on the 12C + 12C reaction rate, we follow the prediction of Perez-Torres et al. (2006) and assume the existence of a single, narrow resonance with ER = 1.5 MeV. Then

Equation (42)

Equation (43)

where 〈σvNR is the nonresonant contribution to the total reaction rate as given in, e.g., Caughlan & Fowler (1988), 〈σvR is the resonant contribution,

Equation (44)

is the resonance strength, J is the total angular momentum of the resonance, ΓC is the entrance channel width, and ΓR ≫ ΓC is the resonance width.

The resonant contribution 〈σvR ∝ (ωγ)R. Using the Breit–Wigner single resonance formula,

Equation (45)

(e.g., Clayton 1983). Evaluating Equation (45) at E = ER,

Equation (46)

where we normalize the resonance width ΓR to that typical of known resonances (see, e.g., Table IV of Aguilera et al. 2006) and the cross section at resonance σ(ER) to the approximate value Perez-Torres et al. (2006) predict. For this work, we adopt (ωγ)R = 3.4 × 10−8 eV as the fiducial resonance strength.

To determine an upper limit for (ωγ)R, we demand that the resonance's contribution to the astrophysical S-factor at a given energy E, SR(E), be less than the experimentally measured value Sexp(E) for all E ≳ 2.1 MeV, the lowest energy probed at the time of this writing. The S-factor for 12C + 12C is

Equation (47)

(Patterson et al. 1969; Clayton 1983). From Equations (45) and (47), we write

Equation (48)

Using Equations (45), (47), and (48), demanding that SR(E) < Sexp(E), and noting that (EER)2 ≫ (ΓR/2)2, we find

Equation (49)

for a resonance at ER = 1.5 MeV. Equation (49) must be satisfied for all E. According to the experimental data, the minimum value of the bracketed term is ≈1 (see, e.g., Figure 4 of Spillane et al. 2007), so

Equation (50)

If ΓR ≈ 100 keV, then our fiducial strength (ωγ)R = 3.4 × 10−8 eV is comparable to the maximum possible strength. ΓR may be much smaller, however; the resonance at 2.14 MeV, the lowest-energy resonance known as of this writing, has a width ΓR < 12 keV (Spillane et al. 2007). Therefore, we set (ωγ)R = 3.4 × 10−7 eV, which is 10 times larger than our fiducial rate, as a reasonable upper limit.

Figure 2 shows the effect a 1.5 MeV resonance has on the reaction rate 〈σv〉. For the fiducial (ωγ)R value, the resonance increases 〈σv〉 by a factor ≳25 at temperatures relevant to superbursts; for the (ωγ)R upper limit, the resonance increases 〈σv〉 by a factor ≳250. These increases are of the order required to reconcile the observationally inferred Σign with that calculated from theoretical models for a specific range of assumed crust thermal conductivities and core neutrino emissivities. In the following section, we compute the superburst Σign with the effect of this resonance.

Figure 2.

Figure 2. Ratio of the total thermally averaged reaction rate 〈σv〉 = 〈σvNR + 〈σvR to the nonresonant contribution 〈σvNR for a hypothetical 1.5 MeV resonance with strength (ωγ)R = 3.4 × 10−8 eV, the fiducial value, as a function of temperature T = T8 × 108 K. The resonance increases 〈σv〉 by a factor ≳ 25 near T8 ≈ 5.

Standard image High-resolution image

5. EFFECTS OF A RESONANCE ON SUPERBURST IGNITION

We use the global linear stability analysis of Cooper & Narayan (2005) to determine the effect a strong resonance in the 12C + 12C system would have on the superburst ignition depth Σign. We assume steady spherical accretion onto a neutron star of mass M = 1.4 M and radius R = 10 km. The accreted matter composition is that of the Sun: the hydrogen mass fraction X = 0.7, helium mass fraction Y = 0.28, and heavy-element mass fraction Z = 0.02. Furthermore, we follow Cumming et al. (2006) and assume the 12C mass fraction XC = 0.2 at the base of the accreted layer.

We make the following two modifications to the model of Cooper & Narayan (2005). (1) Cooper & Narayan (2005) followed Brown (2000) and assumed the energy generated by electron captures, neutron emissions, and pycnonuclear reactions in the crust was distributed uniformly between Σ12 = 6 × 103 and 2 × 105. We now follow Haensel & Zdunik (2008) and distribute the energy according to their Table A.3. (2) Plasma screening reduces the entrance channel width ΓC. Therefore, the plasma screening enhancement for the resonant contribution to the reaction rate includes a correction factor that reduces the overall enhancement (Salpeter & Van Horn 1969; Mitler 1977), although the reduction is only a few percent for the conditions relevant for superbursts (see, e.g., Figure 1 of Cussons et al. 2002). We now use the formalism of Itoh et al. (2003) for the plasma enhancement factors of both the resonant and nonresonant contributions.

The 12C + 12C reaction rate, accretion rate $\dot{\Sigma }$, and ocean temperature profile together determine Σign. The temperature profile is a strong function of the crust's thermal conductivity and core's neutrino emissivity, both of which are poorly constrained. We parametrize these uncertainties by implementing two conductivity and three core neutrino emissivity prescriptions that likely bracket their true values in accreting neutron stars. The thermal conductivity is a decreasing function of the impurity parameter Qimp = 〈Z2〉 − 〈Z2 (Itoh & Kohyama 1993, see also Daligault & Gupta 2009). Schatz et al. (1999) found Qimp ∼ 100 from steady state nucleosynthesis calculations, although subsequent calculations suggest that Qimp should be smaller (Schatz et al. 2003a; Woosley et al. 2004a; Koike et al. 2004; Horowitz et al. 2007, 2009). In addition, fits to the quiescent light curves of KS 1731 − 260 (Shternin et al. 2007; Brown & Cumming 2009) and MXB 1659 − 29 (Brown & Cumming 2009) require that Qimp ∼ 1. Since both observations and molecular dynamics simulations imply that the crust forms an ordered lattice, we adopt Qimp = 3 and 100 as the two bracketing values. The core neutrino emissivity, and thereby the core cooling rate, depends on the unknown ultradense matter equation of state (for reviews, see Yakovlev & Pethick 2004; Page et al. 2006). We consider one "fast" cooling model for which the pion condensate process dominates and two "slow" cooling models for which either the modified Urca or nucleon–nucleon bremsstrahlung process dominates (see, e.g., Table 1 of Page et al. 2006); these roughly correspond to cases "A," "B," and "D" of Cumming et al. (2006, see their Table 2). The respective core temperatures for these models are approximately 3 × 107 K, 3 × 108 K, and 6 × 108 K.

Figure 3 shows the superburst ignition column depth Σign as a function of $\dot{\Sigma }/\dot{\Sigma }_{\mathrm{Edd}}$ for various neutron star models.7 A 1.5 MeV resonance in the 12C + 12C system lowers Σign by a factor ≈2 and ≈4 for the fiducial and maximum (ωγ)R values, respectively; the lowered Σign values are in accord with the observationally inferred values for a range of realistic neutron star model parameters. Therefore, we conclude that (1) a strong resonance may exist at an energy ≈1.5 MeV above the 12C + 12C ground state, and (2) if such a resonance exists, it will mitigate the discrepancy between observationally inferred superburst ignition depths and those calculated from theoretical models.

Figure 3.

Figure 3. Superburst ignition column depth Σign as a function of the Eddington-scaled accretion rate $\dot{\Sigma }/\dot{\Sigma }_{\mathrm{Edd}}$ for various model parameters. Solid (dashed) lines show results for models with impurity parameter Qimp = 3(100). "Pion," "Modified Urca," and "Bremsstrahlung" refer to the core's dominant neutrino emission mechanism. For a given Qimp and neutrino emission mechanism, the three lines show results for a 12C + 12C reaction rate with no resonances (the standard rate), a hypothetical 1.5 MeV resonance with the fiducial strength (ωγ)R = 3.4 × 10−8 eV, and a hypothetical 1.5 MeV resonance with an approximate maximum strength (ωγ)R = 3.4 × 10−7 eV, from top to bottom. The boxes show the inferred Σign and $\dot{\Sigma }$ ranges for the majority of observed superbursts. A 1.5 MeV resonance lowers Σign by a factor ≈2 and ≈4 for the fiducial and maximum (ωγ)R values, respectively.

Standard image High-resolution image

For the low-mass X-ray transient 4U 1608 − 522, which exhibited a superburst, the thermal quiescent luminosity constrains the core temperature to be ≈2.5 × 108 K. Fits to the superburst light curve find an ignition column Σign = (1.5–4.1) × 1012 g cm−2. A resonance at 1.5 MeV could make the ignition temperature over this range as low as (4.1–4.8) × 108 K, which is marginally consistent with the calculated crust temperature at the time of the superburst (Keek et al. 2008).

For the transient KS 1731 − 260, the timescale for the effective temperature to decrease implies Qimp ≲ 1, and the lowest observed effective temperature implies that the core temperature is ≲108 K (Shternin et al. 2007; Brown & Cumming 2009). Under these conditions, the temperature at Σ ≈ 1012 g cm−2 is unlikely to be >3 × 108 K and therefore too cold to match the inferred ignition depth, even if the proposed resonance exists. Recent theoretical calculations of nuclear reactions in the neutron star crust suggest that the pycnonuclear fusion of neutron-rich, low-Z ions such as 24O (Horowitz et al. 2008) or reactions triggered by β-delayed neutron emissions (Gupta et al. 2008) may provide a strong source of heating for the neutron star outer crust. Indeed, fits to the quiescent light curves of KS 1731 − 260 and MXB 1659 − 29 suggest that the heating in the outer crust is larger than can be accounted for from electron captures (Brown & Cumming 2009). Although a survey of neutron star models with this additional heating is outside the scope of this paper, we note that a strong resonance does alleviate the discrepancy in Σign even if it does not entirely resolve it.

6. SUMMARY AND DISCUSSION

In this work, we reexamined the superburst trigger mechanism to address the discrepancy between observationally inferred superburst ignition column depths Σign and those calculated in theoretical models. Motivated by the suggestion of Kuulkers et al. (2002b) and the similarity between inferred ignition column depths from different sources, we first explored the viability of thermally unstable electron captures as the trigger mechanism in Section 2. We found that electron captures are always thermally stable in accreting neutron star oceans; thus electron captures do not trigger superbursts. We then investigated the viability of nuclear fusion reactions other than 12C + 12C. Accretion-induced nuclear reactions deplete ions with Z < 6 at column depths Σ ≪ Σign, whereas ions with Z > 6 fuse at Σ ≫ Σign. We therefore confirmed the proposal (Cumming & Bildsten 2001; Strohmayer & Brown 2002) that 12C + 12C triggers superbursts.

We then examined the 12C + 12C fusion rate in Section 4, noting that superburst model results would be in accord with observations if the true fusion rate were greater than the standard rate by a factor ≳102. Two factors determine the fusion rate: plasma screening effects and the nuclear cross section σ(E). Uncertainties in, and corrections to, the plasma screening enhancement to the reaction rate alter the usual enhancement by a factor <2 and thus cannot resolve the discrepancy between superburst observations and theoretical models. However, uncertainties in σ(E) are much larger; indeed, σ(E) is experimentally unknown at astrophysically relevant energies. We find that a strong resonance in the 12C + 12C system at an energy near 1.5 MeV could increase the fusion rate by a few orders of magnitude at the temperatures relevant to superbursts. Both theoretical optical potential models and extrapolations of existing experimental data suggest that a resonance exists at an energy near 1.5 MeV. If this is true and the resonance strength (ωγ)R is sufficiently large, it could eliminate the discrepancy between observationally inferred superburst ignition column depths and theoretical model results (see Figure 3).

In Section 1, we outlined three fundamental problems that exist with superburst ignition. We address these problems below in the context of our results.

  • 1.  
    The results of all previous superburst models imply that ocean temperatures are too low for 12C ignition at the inferred Σign ≈ 1012 g cm−2. A strong resonance near 1.5 MeV in the 12C + 12C system would decrease the temperature required for ignition at Σign ≈ 1012 g cm−2 from ≈6 × 108 K to ≈5 × 108 K.
  • 2.  
    Heavy-ion fusion hindrance would imply that the standard S-factor overestimates the true 12C + 12C fusion rate. This existence of such hindrance is currently speculative for both 12C + 12C in particular (Jiang et al. 2007) and exothermic fusion reactions in general (Jiang et al. 2008; Stefanini et al. 2008). Furthermore, the effect heavy-ion fusion hindrance has on resonant reactions is unknown. Therefore, it is unclear whether heavy-ion fusion hindrance poses a problem for superburst ignition.
  • 3.  
    The 12C yield from nucleosynthesis models is often lower than that required for a thermal instability. A strong resonance would reduce the minimum 12C abundance required for a superburst, but by only a small amount. Thus, this problem would be attenuated but not resolved.

Our result has implications for 12C + 12C reactions in other contexts, namely Type Ia supernovae and massive stellar evolution. In addition, we have also presented a general prescription for understanding the screening enhancement factor in a multicomponent plasma. We briefly describe each of these topics before concluding with an outlook on future measurements.

6.1. Implications for Type Ia Supernovae and Massive Stellar Evolution

The fusion of 12C is an important stage in the post-main-sequence evolution of a massive star, and it is the reaction that ignites a white dwarf and triggers a thermonuclear (Type Ia) supernova. In both systems, the competition between heating from the 12C + 12C reaction and cooling from neutrino emissions determines ignition. To explore the implications of a resonance in the reaction cross section on these phenomena, we construct ignition curves (Figure 4), defined as epsilonnuc(ρ, T) = epsilonν(ρ, T), for a 12C-16O plasma with XC = 0.5. We compute the neutrino emissivity for the pair, photo, plasma, and bremsstrahlung processes using analytical fitting formulae (Itoh et al. 1996). For epsilonnuc, we use the effective reaction $\mathcal {Q}$-value of 9.0 MeV (Chamulak et al. 2008), which includes heating from both the p- and α- branches and subsequent reactions; the ignition curve is insensitive to the choice of $\mathcal {Q}$. Three curves are plotted in Figure 4 for different choices of the 12C + 12C rate: the standard nonresonant rate (Caughlan & Fowler 1988, dotted line), a resonance at ER = 1.5 MeV with our fiducial strength (ωγ)R = 3.4 × 10−8 eV (solid line), and a resonance at ER = 1.5 MeV with our maximum strength (ωγ)R = 3.4 × 10−7 eV (dashed line).

Figure 4.

Figure 4. Locus in the temperature-density plane where epsilonnuc = epsilonν, which defines the ignition of 12C for stellar burning in massive stars and for thermonuclear (Type Ia) supernovae. The composition is 12C-16O with XC = 0.5. The three curves, from top to bottom, show epsilonnuc computed with the standard rate (Caughlan & Fowler 1988, dotted line), with a resonance at our fiducial strength, (ωγ)R = 3.4 × 108 eV (solid line), and with a resonance at our maximum strength, (ωγ)R = 3.4 × 10−7 eV (dashed line).

Standard image High-resolution image

As is evident from Figure 4, the effect of a resonance at ER = 1.5 MeV is minimal for the ignition of Type Ia supernovae, which are thought to ignite at central densities >109 g cm−3 (see, e.g., Woosley et al. 2004b). It is interesting to speculate that a resonance at a lower energy might shift the ignition curve to lower densities. This would reduce the in situ neutronization of the nickel-peak material synthesized in the explosive burning and the neutronization during the pre-explosive convective burning (Piro & Bildsten 2008; Chamulak et al. 2008); moreover, numerical simulations (Röpke et al. 2006) find that a lower central density reduces the growth of the turbulent flame velocity (because the lower gravitational acceleration decreases the growth rate of the Rayleigh–Taylor instability), which leads to a less vigorous explosion and a decreased production of iron-peak elements. Although lower densities are necessary to avoid overproduction of neutron-rich isotopes such as 54Fe and 58Ni (Woosley 1997; Iwamoto et al. 1999), this may be ameliorated by improved electron capture rates onto pf-shell nuclei (Martínez-Pinedo et al. 2000; Brachwitz et al. 2000). Moreover, there is observational evidence that the majority of supernovae do undergo electron captures in the innermost ≈0.2 M of ejecta (Mazzali et al. 2007). Given the uncertainties in modeling the progenitor evolution, flame ignition, and explosion, and in the dependence of the ignition density on the accretion history of the white dwarf (see Lesaffre et al. 2006, for a recent discussion), we do not think it possible to constrain the existence of such a resonance from observations at this time, but future modeling efforts should clearly allow for this possibility.

Intriguingly, the largest effect is at ρ ≲ 105 g cm−3, which is the region encountered by post-main-sequence massive stars (see Woosley et al. 2002, and references therein). Stellar evolutionary calculations, which include the shock-induced explosive nucleosynthesis, find that a decrease in the 12C + 12C rate leads to enhancements in 26Al and 60Fe abundances (Gasques et al. 2007). Further calculations are needed to determine whether an enhanced 12C + 12C rate would produce interesting changes in nucleosynthesis.

6.2. Ignition in a Multicomponent Plasma

In Section 4.1.1, we showed that heavy spectator ions increase the plasma screening enhancement to the thermonuclear reaction rate via linear mixing rule deviations. To our knowledge, this work is the first to show that spectator ions affect the plasma screening enhancement in the thermonuclear regime. Prior work on the effects such deviations have on the plasma screening enhancement focused on binary ionic mixtures consisting only of reactants and products (e.g., 12C and 24Mg) but no spectator ions (e.g., Ogata et al. 1993; DeWitt et al. 1996), and DeWitt & Slattery (2003) concluded that linear mixing rule deviations always decrease the plasma screening enhancement in binary ionic mixtures (Equation (36)). However, determining the effect of spectator ions requires analyzing a mixture of three or more ions, but previous applications of linear mixing rule deviations in ternary ionic mixtures (e.g., Ogata et al. 1993) focused only on phase diagrams of crystallizing white dwarfs (e.g., Isern et al. 1991; Segretain 1996), not on fusion reactions.

In our analysis, we tacitly assumed that the plasma is uniformly mixed. However, recent molecular dynamics simulations of multicomponent plasmas exhibit clustering of low-Z ions (Wünsch et al. 2008; Horowitz et al. 2009), which may enhance the reaction rate. This is worthy of further investigation.

For completeness, we note that linear mixing rule deviations are much larger for Coulomb solids (DeWitt & Slattery 2003). This has two consequences for accreting neutron stars. (1) Screening enhancements for multicomponent plasmas in the pycnonuclear regime may be orders of magnitude greater than currently thought. This could lower the pressures at which pycnonuclear reactions occur in the crust and thereby heat the ocean to a larger extent. (2) A multicomponent plasma's freezing temperature is much lower than that of a one-component plasma. This possibly explains the results of Horowitz et al. (2007), whose molecular dynamics simulation of a multicomponent plasma froze at Γ ≈ 247 rather than the typical Γ ≈ 175 of a one-component plasma.

6.3. Outlook for Future Measurements

Our conclusions are contingent on the existence of a strong resonance near 1.5 MeV in the 12C + 12C system. As noted in Section 4.2, resonances exist throughout the experimentally studied energy range and are spaced at intervals of ≈0.3 MeV. Therefore, a resonance almost certainly exists sufficiently near 1.5 MeV (i.e., within the Gamow window). We cannot predict with confidence, however, that the resonance strength (ωγ)R is sufficiently large. Indeed, the measured resonances at higher energies typically increase the thermally averaged reaction rate 〈σv〉 by a factor ≲10 over the nonresonant contribution, so the resonance needs to be unusually strong. Resolving this issue requires experimental measurements of the 12C + 12C cross section near the Gamow peak, which requires lab energies of 3 MeV with the ability to measure cross sections at the 0.1 pb level. Measurements at higher energies are probably required to map out the resonance structure between the Gamow peak and the other available measurements of the fusion cross section. Such measurements are possible in the near term with existing laboratories and are certainly within reach of planned underground facilities such as DUSEL (though they will require the larger DUSEL accelerator option; Görres & Wiescher 2006).

We thank Lars Bildsten, Philip Chang, Andrey Chugunov, Richard Cyburt, Daniel Kasen, Hendrik Schatz, Michael Wiescher, Dima Yakovlev, Remco Zegers, and the anonymous referee for their advice and feedback. The Joint Institute for Nuclear Astrophysics (JINA) supported this work under NSF-PFC grant PHY 02-16783. A.W.S. and E.F.B. are supported by NASA/ATFP grant NNX08AG76G, and R.L.C. is supported by the National Science Foundation under grant No. NSF PHY05-51164.

APPENDIX: DERIVATION OF Δ H(0)

In this appendix, we derive ΔH(0), the correction to the plasma screening enhancement factor due to linear mixing rule deviations. Consider a strongly coupled Coulomb plasma consisting of N ≡ ∑iNi ions, where Ni is the number of ions with charge Zi, xiNi/N is the number fraction of species i, and species 1 and 2 are the reactant and product of the reaction, respectively, so that Z2 = 2Z1. DeWitt et al. (1973) found that, neglecting quantum contributions, H(0) equals the difference in the Coulomb free energy before and after the reaction; since a fusion reaction destroys two reactant ions and creates one product ion,

Equation (A1)

Writing Equation (A1) in a more general form,

Equation (A2)

where ΔN2 is the number of products created. In the limit ΔN2N1, N2, Equation (A2) simplifies to

Equation (A3)

Using the general expression for Fex (Equation (34)), we find

Equation (A4)

where the term 2fex1) − fex2) is the well-known result for a plasma obeying the linear mixing rule (e.g., Jancovici 1977), the bracketed term is ΔH(0)/kBT, and we have defined the operator

Equation (A5)

Potekhin et al. (2009) derived an accurate, analytic fitting formula for Δfex (see their Equation (16)). Using their formula for an unpolarized electron background, which is appropriate for a strongly coupled Coulomb liquid, Δfex = Δfex(Γ, 〈Z〉, 〈Z2〉, 〈Z5/2〉), where Γ = ∑ixiΓi and 〈Zk〉 = ∑ixiZki. From Equations (A4) and (A5) of this work and Equations (12), (14), and (16) of Potekhin et al. (2009), we find

Equation (A6)

where

Equation (A7)

Footnotes

  • In this and following expressions, we suppress the scaling with g and evaluate the expressions at g = 2.43 × 1014 cm s−2.

  • We define the Eddington luminosity to be 4πGMces in the frame of a distant observer. This is the largest luminosity observable by such an observer (see Shapiro & Teukolsky 1983). This differs from the definition used by Galloway et al. (2008) by a factor of [1 − 2GM/(Rc2)]−1/2 = 1.3.

  • This statement is not strictly true for heavy-ion fusion reactions such as 12C + 12C. The compound nucleus 24Mg has numerous quasi-stationary states at excitation energies near 15 MeV above the ground state (Endt 1990; Firestone 2007), where Epk lies; thus all reactions are resonant. However, when the mean level spacing of quasi-stationary states DkBT, one computes an average cross section over all resonances, and the reaction rate assumes the nonresonant form (e.g., Cameron 1959; Fowler & Hoyle 1964)

  • The average level spacing of the detected resonances is much greater than that of the quasi-stationary states in the compound nucleus 24Mg. Thus the observed resonances are not ordinary compound nuclear states. As Almqvist et al. (1960) first suggested, the resonances are probably quasi-molecular doorway states in the 12C + 12C system (e.g., Betts & Wuosmaa 1997).

  • The critical $\dot{\Sigma }$ below which 12C burns stably calculated in our global stability analysis is lower than that calculated in the one-zone model of Cumming et al. (2006, compare to their Figure 15). The reason is simple: following Cumming & Bildsten (2001), they demand that the characteristic lifetime of a 12C ion XC/rnuc > tacc. However, rnuc depends exponentially on the density due to plasma screening (Section 4.1), so 12C burning occurs in a narrow column depth range, much like the electron captures discussed in Section 2.3. Thus, the proper criterion is XC/rnuc > taccΔΣ/Σ, where ΔΣ/Σ is similar to the expression given in Equation (10). This proper criterion gives a lower critical $\dot{\Sigma }$, in accord with our results.

Please wait… references are loading.
10.1088/0004-637X/702/1/660