Abstract

Spectroscopy is used to measure the elemental abundances in the outer layers of the Sun, whereas helioseismology probes the interior. It is well known that current spectroscopic determinations of the chemical composition are starkly at odds with the metallicity implied by helioseismology. We investigate whether the discrepancy may be due to conversion of photons to a new light boson in the solar photosphere. We examine the impact of particles with axion-like interactions with the photon on the inferred photospheric abundances, showing that resonant axion–photon conversion is not possible in the region of the solar atmosphere in which line formation occurs. Although non-resonant conversion in the line-forming regions can in principle impact derived abundances, constraints from axion–photon conversion experiments rule out the couplings necessary for these effects to be detectable. We show that this extends to hidden photons and chameleons (which would exhibit similar phenomenological behaviour), ruling out known theories of new light bosons as photospheric solutions to the solar abundance problem.

INTRODUCTION

Agreement between predictions of solar interior models and helioseismological measurements has deteriorated in recent years, due to a downwards revision in standard solar abundances. Recent photospheric analyses (Allende Prieto, Lambert & Asplund 2001, 2002; Asplund et al. 2004, 2005b, 2009; Asplund, Grevesse & Sauval 2005a; Scott et al. 2006, 2009; Meléndez & Asplund 2008) have indicated that the solar metallicity is about 20 per cent lower than previously thought (Grevesse & Sauval 1998). In particular, the current reference photospheric abundances (Asplund et al. 2009) of the important elements C, N and O are smaller than the previous reference values (Grevesse & Sauval 1998) by 0.09, 0.09 and 0.14 dex, respectively. Standard solar models computed with the revised surface metal abundances show large discrepancies with the depth of the convection zone, the surface helium abundance and the sound speed profile inferred from helioseismology (Basu & Antia 2004, 2008; Bahcall et al. 2005a; Bahcall, Serenelli & Basu 2006; Yang & Bi 2007; Serenelli et al. 2009). This has been variously referred to as ‘the solar model(ling) problem’ (Drake & Testa 2005; Asplund 2008; Morel & Butler 2008), ‘the solar oxygen crisis’ (Ayres, Plymate & Keller 2006; Socas-Navarro & Norton 2007; Ayres 2008) or ‘the solar abundance problem’ (Guzik & Mussack 2010; Serenelli, Haxton & Peña-Garay 2011).

The most common approaches to the solar abundance problem to date have been to recheck the photospheric abundances (Drake & Testa 2005; Ayres et al. 2006; Ayres 2008; Caffau et al. 2008, 2010; Centeno & Socas-Navarro 2008; Morel & Butler 2008), or to modify solar models in an attempt to accommodate both the measured abundances and data from helioseismology. Efforts in the latter category have included enhanced diffusion (Guzik, Watson & Cox 2005), late-stage accretion of low-metallicity gas (Guzik et al. 2005; Castro, Vauclair & Richard 2007; Guzik & Mussack 2010; Serenelli et al. 2011), modifications of opacities (Badnell et al. 2005; Bahcall, Serenelli & Basu 2005b; Christensen-Dalsgaard et al. 2009), internal gravity waves (Arnett, Meakin & Young 2005; Charbonnel & Talon 2005) and enhanced energy transport in the solar core due to gravitational capture and scattering of dark matter (Cumberbatch et al. 2010; Frandsen & Sarkar 2010; Taoso et al. 2010). Despite nearly a decade of effort (and a few errant claims to the contrary), none of these approaches has proven successful in solving the problem.

Here, we propose an alternative approach: examining effects of new physics on the measurement of solar photospheric abundances. Specifically, we investigate the impact on derived elemental abundances of mixing between photons and new light bosons, such as axions, in the line-forming regions of the photosphere.1

The 5770 K thermal spectrum of the Sun is formed around optical depth τ = 1, which defines the location of the solar ‘surface’. For a height z above the surface,
\begin{equation} \tau _\lambda (z) \equiv \int _z^\infty \alpha _\lambda (z^{\prime })\, \mathrm{d}z^{\prime }. \end{equation}
(1)
Here, |$\alpha _\lambda (z) \equiv \kappa _\lambda ^{\rm c}(z) \rho (z)$| is the linear attenuation coefficient at wavelength λ, where |$\kappa _\lambda ^{\rm c}$| is the continuum opacity and ρ is the local gas density. Solar abundances are inferred from absorption lines. These are formed as atomic or molecular bound–bound transitions increase the opacity in small wavelength ranges around the transitions. The higher opacity in these spectral lines means that the light at those wavelengths escapes from layers above the continuum-forming layers (the photosphere). As the gas temperature decreases with height, the line-centre photons escape from cooler layers than the continuum photons, giving rise to absorption lines.

The equivalent width of a spectral line is defined as the width in wavelength space of a rectangle having the same area as the integrated region between the line and the continuum (where the height of the rectangle is given by the continuum level). Equivalent width is a relative measure of line strength, and depends not only on the abundance of a given element, but also on the attenuation of the total continuum as it passes through the line-forming region. For this reason, both line and continuum fluxes must be carefully calculated with radiative transfer simulations before elemental abundances can be inferred. If additional plasma effects in the photosphere enhance continuum photon attenuation via oscillation into light bosons such as axions, then the equivalent width of absorption lines used to calculate elemental abundances would be suppressed. By combining the axion conversion effect with earlier radiative transfer calculations, we will show how this could occur in principle, and assess whether this effect could arise from existing particle theories in such a way as to solve the solar abundance problem.

Historically, the axion field arose as a candidate solution to the strong CP problem of QCD. Although the term
\begin{equation} \tilde{\Theta }_3 G_{\mu \nu }\tilde{G}^{\mu \nu } \end{equation}
(2)
is not forbidden in any way in the QCD Lagrangian, it seems to be absent from the standard model (SM) of particle physics. Here, Gμν is the gluon field strength, and its dual is |$\tilde{G}^{\mu \nu } = \epsilon ^{\alpha \beta \mu \nu } G_{\alpha \beta }$|⁠. The coupling constant |$\tilde{\Theta }_3$| is experimentally consistent with zero, so some sort of dynamical method is required in order to force it to be small without fine-tuning. The Peccei–Quinn (PQ; Peccei & Quinn 1977) mechanism introduces a broken U(1) symmetry that naturally yields this cancellation, with the added effect of producing a new light, propagating, pseudo-scalar degree of freedom: the axion.
The proposed existence of the axion field has given rise to a rich variety of phenomenological predictions. These include axion production in the dense cores of stars (see Raffelt 1990 and references therein), ‘light shining through wall’ effects (Van Bibber et al. 1987) and many theories of axionic dark matter (Duffy & van Bibber 2009, and references therein). Many of these effects are due to the effective coupling |$\mathcal {L}_{a\gamma \gamma }$| that must arise between the PQ axion field a and the photon gauge field strength Fμν:
\begin{equation} \mathcal {L}_{a\gamma \gamma } =\frac{g_{a \gamma }}{4}a F_{\mu \nu } \tilde{F}^{\mu \nu } = -g_{a \gamma }a \boldsymbol {E} \cdot \boldsymbol {B}\, \end{equation}
(3)
with coupling constant gaγ. It is this axion–photon–photon vertex that allows axions to convert into photons in the presence of electric (⁠|$\boldsymbol {E}$|⁠) or magnetic (⁠|$\boldsymbol {B}$|⁠) fields, and vice versa. The mass and coupling of a PQ axion must be very small (with ma∝1/gaγ), and currently remain (mostly) far below the threshold of experimental searches. A rough lower limit of 10−6 eV can but put on the axion mass, in order to avoid overclosing the universe, and an upper limit of ma ≲ 10−3 eV ensures that stellar evolution is not significantly modified (Asztalos et al. 2006).

The PQ mechanism in QCD is not the only possible origin of such a particle, however. Axion-like particles (ALPs) are a ubiquitous element in string theory compactifications (Svrcek & Witten 2006; Dasgupta, Firouzjahi & Gwyn 2008), in which the moduli describing the compact extra dimensions can appear to observers as internal U(1) gauge degrees of freedom. More generally, U(1) symmetries appear in many high-energy extensions of the SM, giving rise to exotic particles with interactions similar to the axion. Like axions, ALPs interact with photons via a term of the form as in equation (3). Unlike axions, they do not necessarily solve the strong CP problem, nor require such small couplings or ma ∝ 1/gaγ.

We begin by showing in Section 2 how photon–ALP oscillation can affect the linear attenuation coefficient – and hence the equivalent width of an absorption line – in the solar atmosphere at optical depths τ < 1. We demonstrate in Section 3 that with a large enough coupling gaγ, a low-mass particle with a term like equation (3) could affect the solar atmosphere enough to modify absorption lines in an observable way. As we discuss in Section 4, however, the coupling strength required to produce such an effect is much larger than the coupling allowed by current experimental bounds on standard ALPs. We go on to show that this rules out not only ALPs as photospheric solutions to the solar abundance problem, but also hidden photons and chameleons. We conclude in Section 5 that unless some other light boson with an axion-like or similar kinetic-mixing-type coupling to the photon can be found that somehow evades the existing experimental constraints, the solar abundance problem cannot be solved by light bosons in the photosphere.

THEORY

Changing inferred abundances

We first consider the standard case, without ALPs. On the linear portion of the curve of growth, the equivalent width of an absorption line Wλ is approximately proportional to the local ratio of the line-centre (⁠|$\kappa _\lambda ^\mathrm{l}$|⁠) to continuum (⁠|$\kappa _\lambda ^\mathrm{c}$|⁠) opacities at the height of line formation (Rutten 2003):
\begin{equation} \!W_\lambda \varpropto \frac{\kappa _\lambda ^\mathrm{l}}{\kappa _\lambda ^\mathrm{c}}\equiv \eta _\lambda . \end{equation}
(4)
The abundance dependence of Wλ is encoded in ηλ by way of |$\kappa _\lambda ^\mathrm{l}$|⁠, which is itself proportional to the number of absorbers along the line of sight. The number of absorbers is given by the number of atoms with electrons in the lower level of the transition, which is some known (temperature- and density-dependent) fraction of the total abundance of the element in question, ϵ.
In the presence of ALPs, the local continuum opacity receives an additional effective contribution |$\kappa _\lambda ^\mathrm{a}$| from photon–ALP conversion. In this case,
\begin{equation} \!W_\lambda \varpropto \frac{\kappa _\lambda ^\mathrm{l}}{\kappa _\lambda ^\mathrm{c}+\kappa _\lambda ^\mathrm{a}}= \frac{\eta _\lambda }{1 + \eta _\lambda ^\mathrm{a}}, \end{equation}
(5)
where we have defined the ‘axionic extinction ratio’ |$\eta _\lambda ^\mathrm{a}\equiv \kappa _\lambda ^\mathrm{a}/\kappa _\lambda ^\mathrm{c}$| in analogy with the line extinction ratio ηλ. The abundance dependence of equation (5) is identical to that of equation (4), i.e. linear in ϵ, with the same constant of proportionality. However, given that Wλ is known from observation, analysis using either equation (4) or (5) is constrained to return the same answer for Wλ. Denoting the abundance derived by ignoring ALPs (equation 4) as ϵstd, and that including the effects of ALPs (equation 5) as ϵcorr, we have
\begin{equation} \!\frac{W_\lambda \ ({\rm equation }\,\,4)}{W_\lambda \ ({\rm equation }\,\,5)} = \frac{\epsilon _\mathrm{std}}{\epsilon _\mathrm{corr}/(1 + \eta _\lambda ^\mathrm{a})}=1. \end{equation}
(6)
The abundance correction to be applied to the standard result in order to obtain the ALP-corrected abundance is thus
\begin{equation} \!\Delta \log \epsilon \equiv \log \epsilon _\mathrm{corr} - \log \epsilon _\mathrm{std} = \log (1+\eta _\lambda ^\mathrm{a}). \end{equation}
(7)
With equation (7) in hand, the remaining task is then to calculate |$\eta _\lambda ^\mathrm{a}$| in the line-forming regions of the photosphere.

Photon–ALP conversion in the solar atmosphere

Regardless of the precise ALP model, the presence of a coupling term with the photon (equation 3) implies that under the right conditions, a thermal photon in the photosphere may oscillate into an ALP and vice versa. This occurs in the presence of an external photon field: either a bulk magnetic field, or the oscillating fields of the stellar plasma. Only the former effect is relevant in the region of interest for line formation, as plasma field-induced transition rates (equation 9) turn out to be extremely small in the photosphere. In order to arrive at the quantity |$\eta ^a_\lambda$|⁠, we first compute the linear attenuation coefficient |$\alpha ^a_\lambda$|⁠, related to the ALP opacity by
\begin{equation} \alpha ^a_\lambda \equiv \kappa _\lambda ^a \rho. \end{equation}
(8)

Conversion in plasma fields

The transition from photon γ to ALP a induced by the oscillating electric fields in a plasma occurs mainly from resonant Primakoff conversion. Neglecting recoils, the transition rate is given by (Raffelt 1988, 2008)
\begin{equation} \alpha _{\lambda ,\mathrm{pl}}^a = \frac{g_{a \gamma }^2 T \kappa _s^2}{32\pi }\left[\left(1+\frac{\kappa _s^2}{4\omega ^2} \right) \ln \left(1+\frac{4\omega ^2}{\kappa _s^2}\right) \right]\!, \end{equation}
(9)
where κs is the screening scale in the Debye–Hückel approximation:
\begin{equation} \kappa _s^2 = \frac{4\pi \alpha _{\rm EM}}{T}\left(n_{\rm e} + \sum _{\rm nuclei}Z_{\rm i}^2n_{\rm i}\right)\!, \end{equation}
(10)
where αEM is the fine-structure or electromagnetic coupling constant, Zi is the charge of the ith ion and the number densities ne, ni of electrons and ions, respectively, are found by solving the equation of state. This effect is highly suppressed at the relatively low charge densities of the solar atmosphere, and contributes a negligible amount at optical depths of interest to us.

Conversion in a bulk magnetic field

Conversion in a bulk magnetic field can be much more significant than in the plasma field. The probability that a photon polarized in the direction of the |$\boldsymbol {B}$| field will have converted to an ALP after travelling a distance δz ≡ z − z0 from height z0 to height z is (Raffelt & Stodolsky 1988; Mirizzi, Raffelt & Serpico 2008)
\begin{equation} P_0(\delta z) = 4\left(\frac{\Delta _{a \gamma }}{\Delta _{\rm osc}}\right)^2 \sin ^2(\Delta _{\rm osc}\delta z/2), \end{equation}
(11)
where
\begin{eqnarray} &&\!\!\!\!\Delta _{\rm osc}^2 = (\Delta _{||}-\Delta _a)^2 + 4\Delta _{a \gamma }^2, \nonumber\\ &&\!\!\!\!\Delta _{||} = \Delta _{||}^{\rm {vac} }+ \Delta _{||}^{\rm {gas}} + \Delta _{\rm pl}\nonumber \\ &&\!\!\!\!\Delta _a = -\frac{m_a^2}{2\omega } \nonumber \\ &&\!\!\!\!\Delta _{a \gamma } = g_{a \gamma }|\boldsymbol {B}_T|/2, \nonumber \\ &&\!\!\!\!\Delta _{\rm pl} = -\frac{\omega _{\rm pl}^2}{2\omega }. \end{eqnarray}
(12)
|$\boldsymbol {B}_T$| is the component of the magnetic field transverse to the direction in which the photon propagates, ω = 2πc/λ is the frequency of the incoming photon and the plasma frequency is ωpl = 4παEMne/me. Due to the |$\boldsymbol E\cdot \boldsymbol B$| form of the interaction (3), only the photons with polarization parallel to the magnetic field are susceptible to conversion; this is the meaning of the || subscript. |$\Delta _{||}^{\rm {vac}} = \omega (n_{||}^{\mathrm{vac}}-1)$|⁠, where |$n_{||}^{\mathrm{vac}} \sim 1 + (\alpha _{\rm EM} B/4\pi B_{\mathrm{c}})$| is the ‘vacuum’ contribution to the refractive index induced by the magnetic field. The critical field |$B_c = m_e^2/e \sim 10^{12}\,\mathrm{eV}^2 \sim 4 \times 10^9$| T is much larger than the fields B ∼ 0.1 T that we consider in this work; |$\Delta _{||}^{\rm {vac}}$| is therefore negligible for all cases of interest here. Similarly, |$\Delta _{||}^{\rm {gas}} = \omega (n_{||}^{\rm {gas}} -1)$|⁠, where |$n_{||}^{\rm {gas}}$| is the contribution to the refractive index by solar mixture gas. The main contributors are Rayleigh scattering by neutral atomic and molecular hydrogen as well scattering from free electrons. |$n_{||}^{\rm {gas}}$| for bluer wavelengths is dominated by neutral scattering, whereas free electron scattering becomes more important for larger values of λ. A relativistic calculation of the static polarizability of hydrogen (Bartlett & Power 1969) gives |$\alpha ^0_{\mathrm{H}{\small I}} = 6.667\,9437 \times 10^{-25} \, \mathrm{cm}^3$|⁠. A good approximation is
\begin{equation} n_{||}^{\rm {gas}} = \sqrt{1 + 4\pi \left(n^{\phantom{0}}_{\mathrm{H}\, {\small I}}\alpha ^0_{\mathrm{H}\,{\small I}} + n_{\rm e}\alpha _{\rm e}\right)}, \end{equation}
(13)
where |$n^{\phantom{0}}_{\mathrm{H}\small {I}}$| and ne are, respectively, the number densities of neutral hydrogen atoms and free electrons, and αe = (e/ω)2/me. This is valid in the low densities of the solar atmosphere and as long as we are sufficiently far from the Lyman α line. The red curve in Fig. 1 shows this contribution via |$(\omega \Delta _{||}^{\rm {gas}})^{1/2}$|⁠, where we have also included the subdominant effects of He i and H2 scattering, as well as H i line-scattering.
Figure 1.

Green: average plasma frequency ωpl (eV) as a function of optical depth at 500 nm (bottom axis) and averaged height (upper axis) in a convection simulation of the Solar atmosphere. Red: gas contribution to Δ||, plotted as |$(\omega \Delta _{||}^{\rm gas})^{1/2}$| at photon energies corresponding to λ = 300 nm (dashed – primarily neutral hydrogen) and 900 nm (solid – primarily electrons). When the difference between the plasma and gas quantities (thick blue lines) is positive, |$\omega _{\rm pl}^2 - \omega \Delta _{||}^{\mathrm{gas}} > 0$|⁠, resonant enhancement may occur. The value of this difference at a given height corresponds to the ALP mass ma for which resonant conversion takes place.

Equation (11) describes a Lorentzian (Breit–Wigner) distribution in Δa ∝ 1/ω modulated by the sin 2 term, with resonance at the plasma frequency and width ΔaγB. When Δ|| − Δa goes to zero, photon conversion to ALPs is enhanced. We later show that for ma ∼ 10−5–10−3 eV, this has the potential to occur in the region of the solar atmosphere where line formation occurs. The plasma frequency ωpl increases quickly with increasing optical depth. Given the |$\omega _{\rm pl}^{-4}$| suppression of equation (11), this means that conversion is highly suppressed at depths below the resonance depth hres. There are thus three qualitatively different regions in the solar atmosphere: (1) below hres, the denominator of P0 is dominated by the plasma term |$\Delta _{\rm pl}^2 \propto \omega _{\rm pl}^{4}$|⁠, ensuring that very little conversion occurs below the resonance height; (2) at hres, Δpl ≃ Δa and resonance occurs: conversion is enhanced and P0 → 1 for a small region; and (3) above hres, P0 becomes approximately constant as the plasma frequency falls and the constant term Δa essentially dominates Δosc, providing moderate conversion rates. The existence of region (2) is contingent on the plasma effects being larger than the contribution from neutral gas; otherwise, Δ|| and −Δa are always positive, prohibiting cancellation.

We point out that many references (e.g. Raffelt 1988; Mirizzi et al. 2008) do not include |$\Delta _{||}^{\rm {gas}}$|⁠. This is because in situations such as hot stellar interiors or the interstellar medium, the ionization fraction is so large that |$\Delta _{||}^{\rm {gas}}$| is always subdominant to Δpl. Indeed, the intermediate temperatures encountered in the solar atmosphere make it one of the few places where both the gas refractive index and plasma frequency contribute to similar degrees. Fig. 1 illustrates these effects, showing that resonant conversion may only occur below log τ ∼ 0 (i.e. at larger optical depths). The blue lines show the ALP mass at which resonant conversion may occur for two different photon wavelengths.

We now compute the effect of photon–ALP conversion on the photon intensity for two complementary magnetic field configurations, starting with the case where the size of the magnetic domains s is larger than the scaleheight over which line formation occurs. After propagation over a distance δz, photon loss to photon–axion conversion gives the change in intensity (Mirizzi et al. 2008):
\begin{equation} I_\gamma (z) = \left(1-\frac{1}{2}P_0(\delta z)\right)I_\gamma (z_0). \end{equation}
(14)
This assumes that the change of the photon intensity due to radiative transfer is small over the distance δz and does not feedback on the calculation of the conversion rate. The linear attenuation coefficient due to ALP conversion is then
\begin{equation} -\frac{I_\gamma ^{\prime }(z)}{I_\gamma (z)} \equiv \alpha _\lambda ^a = \frac{\Delta _{\rm osc}\sin (\Delta _{\rm osc}\delta z)}{\frac{\Delta _{\rm osc}^2}{\Delta _{a \gamma }^2}-1 + \cos (\Delta _{\rm osc}\delta z)}, \end{equation}
(15)
where the prime symbol denotes differentiation with respect to z. When computing the full radiation transfer, (15) must then be added to the linear attenuation due to opacity.
For a line with some mean optical depth of formation τ, the range of heights over which |$\alpha _\lambda ^a$| will contribute to the line profile is approximately given by the scaleheight Hκ with which the continuum opacity changes by a factor of e,
\begin{equation} H_{\kappa }\equiv \left(\frac{1}{\kappa ^{\rm c}_\lambda }\frac{\mathrm{d}}{\mathrm{d}z}\kappa ^{\rm c}_\lambda \right)^{-1}. \end{equation}
(16)
This simply states the fact that to a rough approximation, the width of a line's contribution function is given by the opacity scaleheight.2 The relevant contribution of |$\alpha _\lambda ^a$| to the continuum opacity for any line formed at depth τ can then be estimated by averaging |$\alpha _\lambda ^a$| over Hκ
\begin{equation} \langle \alpha _\lambda ^a \rangle = \frac{1}{H_{\kappa }}\int _1^{x(H_{\kappa })}\frac{ \mathrm{d}x}{\left(\frac{\Delta _{\rm osc}}{\Delta _{a \gamma }}\right)^2-1+x}, \end{equation}
(17)
where xz) = cos (Δoscδz). Integrating gives
\begin{equation} \langle \alpha _\lambda ^a \rangle = -\frac{1}{H_{\kappa }} \ln \left\lbrace 1-\left(\frac{\Delta _{a \gamma }}{\Delta _{\rm osc}}\right)^2\left[1-\cos (\Delta _{\rm osc}H_{\kappa })\right] \right\rbrace. \end{equation}
(18)
Away from resonance, Δosc ≫ Δaγ and the log may be Taylor-expanded about 1, giving
\begin{equation} \langle \alpha _\lambda ^a \rangle = \frac{1}{H_{\kappa }}\left(\frac{\Delta _{a \gamma }}{\Delta _{\rm osc}}\right)^2\left[1-\cos (\Delta _{\rm osc}H_{\kappa })\right]. \end{equation}
(19)
Because we are interested in the resonance region as well, we will only make use of the full expression (equation 18).
If photons instead propagate through n domains of length s with equal magnetic field strength but random orientation (such that δz = ns), their intensity is reduced to (Mirizzi et al. 2008)
\begin{eqnarray} \left[\begin{array}{c}I_\gamma (z) \\ I_a(z) \end{array} \right] = \left[\begin{array}{c c}1-\frac{1}{2}P_0 &\quad P_0 \\ \frac{1}{2}P_0 &\quad 1-P_0 \end{array} \right]^n \left[\begin{array}{c}I_\gamma (z_0) \\ I_a(z_0) \end{array} \right]\!. \end{eqnarray}
(20)
Here, Ia refers to the intensity of the ALP field and P0 = P0(s). If the ALP population is initially zero – we return to the validity of this assumption below – the photon intensity becomes
\begin{equation} I_\gamma (z) = \frac{1}{3}\left[2+ \left(1-\frac{3}{2}P_0\right)^n\right]I_\gamma (z_0). \end{equation}
(21)
If the opacity scaleheight Hκ (the propagation length of interest) is much larger than the magnetic domain size s, we can use the large n limit of equation (21) (Mirizzi et al. 2008)
\begin{equation} P_{\gamma \rightarrow a}(\delta z) = \frac{1}{3}\left[1-\exp {\left(-\frac{3P_0(s)\delta z}{2s}\right)}\right]\!\!, \end{equation}
(22)
\begin{equation} I_\gamma (z) = \left[1-P_{\gamma \rightarrow a}(\delta z)\right]I_\gamma (z_0). \end{equation}
(23)
This is valid for a photon passing through a large number δz/s of magnetic domains of random orientation, but with identical field amplitudes, B. This serves the same function as the averaging procedure (equation 17). Note that P0 is again evaluated at s, the size of the magnetic domains. The linear attenuation coefficient is then
\begin{equation} \alpha ^a_\lambda =-\frac{I^{\prime }_\gamma (z)}{I_\gamma (z)} = \frac{P_0}{2s}\frac{3}{1+2\,{\rm exp}\left(\frac{3P_0\delta z}{2s}\right)}. \end{equation}
(24)
Away from the resonance region, terms that are higher order in P0 may once more be dropped, yielding
\begin{equation} \alpha ^a_\lambda = \frac{P_0}{2s} + O(P_0^2) \end{equation}
(25)
\begin{equation} \simeq \frac{1}{s}\left(\frac{\Delta _{a \gamma }}{\Delta _{\rm osc}}\right)^2 \left[1 - \cos (\Delta _{\rm osc}s)\right]. \end{equation}
(26)
We see that this is just equation (19), but with the scaleheight Hκ replaced by the magnetic domain size s. For our calculations in Section 3, we use the full expressions (equations 18 and 24) rather than the small P0 approximations.
A final comment should be made on the topic of ALP–photon reconversion. Because this effect is implicit in the definition of P0 (equation 11), it is accounted for in the photon-loss expressions equations (18) and (24). However, we have assumed that at every height the initial ALP population is zero, i.e. Ia(z0) = 0. Although it is not true in general that no axions will reach a given layer from any of the layers below, this is in fact a good approximation for our purposes. At height z0, the ‘initial’ ALP population is given to a good approximation by the conditions in the layer below, so |$I_a(z_0) = \frac{1}{2}P_0(\delta z)I_\gamma (z_0-\delta z)$|⁠. Therefore, equation (20) dictates that the reconversion rate of the initial axions into ‘new’ photons is
\begin{equation} I_{\gamma ,\mathrm{rec}}(z) = \frac{1}{2}P_0(\delta z)^2 I_{\gamma }(z_0-\delta z), \end{equation}
(27)
where Iγ, rec is the intensity of reconverted photons only. As P0 ≪ 1 everywhere but the resonance region, this quantity should be smaller than the photon-loss rate as long as the scaleheight Hκ (and therefore δz) is larger than the width δhres of the resonance region. The latter may be defined as the height over which the prefactor 4(Δaγosc)2 in P0 (11) is O(0.1) or greater. For λ = 900 nm and ma = 10−4 eV, extreme cases that we will consider in Section 3 are gaγ = 0.1 GeV−1, leading to δhres ≃ 30 km; and gaγ = 10−7 GeV−1, giving δhres ≃ 30 m. In all cases, this is smaller than the opacity scaleheight Hκ ≃ 100 km.

EFFECTS ON SOLAR ABUNDANCES

To determine the impact of photon–ALP conversion in the photosphere on solar abundances, we calculated the expected correction factor Δlog ϵ (equation 7) using the 3D hydrodynamic solar model atmosphere from Asplund et al. (2009). We computed the ALP-induced opacity |$\kappa ^a_\lambda$| at wavelengths from 300 to 1000 nm, at each point in an (x, y, z) grid with resolution 50 × 50 × 500, sampled from the atmosphere simulation. Our extracted grid covers 6 × 6 Mm, in the horizontal directions, and optical depths in the range −3 ≤ log τ ≤ 0.5, in which the majority of line formation occurs and below which conversion is negligible. We did this for 90 snapshots from the simulation, corresponding to 45 min in solar time before averaging. All averages presented here are carried out over time and over the undulating surfaces of particular optical depths to give averages on the τ500-scale. We averaged the local values of |$\kappa ^a_\lambda$| over (x, y, t) slices of common optical depth at 500 nm, to arrive at expected abundance corrections as a function of wavelength and formation height of arbitrary spectral lines. Spatial and temporal changes in line profiles due to convective motions occur over the length-scales and lifetimes of convective granules, which are approximately 1 Mm and 10 min for the Sun (Stein & Nordlund 1998; Asplund 2005). As these scales are significantly smaller than the extent of the simulation, our averaged values should accurately represent the expected effect of ALPs on the mean solar spectrum.

The continuum opacities include bound–free absorption by the first three ionization stages of 15 of the most abundant elements, as computed by the Opacity Project and the Iron Project (Badnell et al. 2005). The absorption by the anions of H, He, C, N and O is also included, as well as that by the molecules H2, H|$_2^+$|⁠, H|$_2^-$|⁠, OH, CH, H2O and CO and the pressure induced absorption of transient (H i+H i)-, (H i+He i)-, (H2+H i)-, (H2+He i)- and (H2+H2) pairs. A complete list with detailed references was presented by Hayek et al. (2010). The opacity in the 380–1644 nm range is dominated by the bound–free absorption by H, by free–free absorption of H i and H further into the infrared and by bound–free absorption by H i and various metals further into the ultraviolet.

Depending on initial conditions and simulation details, magnetohydrodynamic models of the photosphere predict vertical magnetic field amplitudes in the quiet Sun up to 0.1 T, with domains ranging from a few hundred to 1000 km in diameter at line-forming heights (Stein & Nordlund 2006; Cheung & Cameron 2012; Moll, Cameron & Schussler 2012). This means that equation (18) is valid for all our calculations; if s were equal to or smaller than the scaleheight Hκ, equation (24) would be needed instead. The horizontal component of the magnetic field can furthermore be two to five times larger than the vertical component (Steiner et al. 2009). We therefore take 〈B〉 = 0.1 T, but caution that some studies (including Steiner et al. 2009) favour a more modest 0.001–0.01 T.

Using a coupling that is only mildly in tension with current constraints (see Section 4.1), gaγ = 10−7 GeV−1, we find that resonant photon–axion conversion has very little impact on the effective opacity: Δlog ϵ < 10−11. This is many orders of magnitude below what is required to solve the solar abundance problem. We illustrate the impact of resonant photon–axion conversion on the effective line opacity in Fig. 2 with ma = 10−4 eV and a very large value of gaγ = 0.1 GeV−1, giving resonant conversion at a mean optical depth of log 〈τ〉 ∼ 0.2. We also show the contribution functions for the 630.0 nm [O i] and 872.7 nm [C i] forbidden lines (reproduced from 3D line formation calculations presented by Caffau et al. (2008) and Caffau et al. (2010), respectively), indicating the contributions from different atmospheric heights to the formation of these lines. These two lines are important contributors to the determination of oxygen and carbon abundances (Allende Prieto et al. 2001, 2002; Asplund et al. 2004, 2005b, 2009). Due to the |$\Delta _{||}^{\rm {gas}}$| contribution, which may be seen as a suppression of the effective photon mass by the index of refraction of neutral hydrogen, resonant conversion within most of the line-forming region is disallowed. We note the sharp rise at low optical depths is because of the sharp drop in the opacity at these heights: Fig. 2(b) shows the relative change in line strength, reflecting the decrease in continuum opacity; the absolute change at optical depths less than ∼−2 has no observable effect.

Figure 2.

The effect of resonant photon–axion conversion on the effective line widths. Left: abundance correction effect in the solar atmosphere due to particles with an axion-like coupling, as a function of optical depth τ and wavelength. Here, ma = 10−4 eV and we have set gaγ = 0.1 GeV−1, to illustrate the effect, although the allowed value of gaγ by observational searches is much lower. Horizontal bars indicate the central 68 per cent intervals of the line formation contribution functions for the 630.0 nm [O i] (green) and 872.7 nm [C i] (black) forbidden neutral atomic lines. Right: constant-wavelength slices through the left-hand panel for (solid lines) λ = 400 nm (magenta), 630 nm (green) and 875 nm (black). The peaks are due to resonant photon–ALP conversion near optical depth log 〈τ〉 = 0. The green and black dashed lines represent the contribution functions dP/dlog τ to the 630.0 nm [O i] and 872.7 nm [C i] absorption lines, respectively. In both cases, ∫dP = 1 with τ being the monochromatic optical depth at the central wavelength of the respective absorption line (rather than τ500).

If the coupling gaγ is pushed to even larger values gaγ > 10 GeV−1, non-resonant conversion can provide corrections of Δlog ϵ > 0.2 dex in the centre of the region of line formation. We illustrate this in Fig. 3. Taken naively, the correction factors we present here could offer at least a partial solution to the solar abundance problem, as abundances from all lines would increase relative to their expected values. This example combination of ma and gaγ would induce the necessary 0.1–0.2 dex change in abundances from some common C and O indicators; similar effects would be expected with other lines used for abundance determination, as they typically form at similar optical depths. The exact amount by which the abundance inferred from each line and species would rise would depend on line choice, line-to-line differences in wavelengths and formation depths, the magnetic field structure and so on.

Figure 3.

The effect of photon–axion conversion above the resonance region, with a coupling large enough to solve the solar abundance problem. ma = 1 × 10−4 eV as in Fig. 2, but gaγ = 10 GeV−1. The solid lines on right-hand figure are for λ = 400 nm (magenta), 630 nm (green) and 875 nm (black); the dashed lines represent the O i and C i contribution functions (see Fig. 2 caption). The conversion probability remains constant throughout the line-forming region of the atmosphere, and the rise in Δlog ϵ is due to the falling bulk opacity with height. We note that the extremely large coupling required is ruled out by several orders of magnitude.

However, to this point we have not carefully considered the viability of the couplings required, in view of independent constraints from solar core or laboratory searches for axions and ALPs. In the following section, we discuss those bounds and the possibility of achieving this effect with real particle models, showing that any observable effect is in fact already ruled out by existing bounds on ALPs.

DISCUSSION

ALPs

There are three types of photon–ALP conversions that can be probed in order to constrain gaγ and ma.3

  • Conversion in hot dense cores of stars should produce X-ray energy ALPs. Observations of horizontal branch stars constrain the rate of energy loss during their He-burning phase (Raffelt 1999). The Sun itself would lose significant amounts of energy from its core if standard ALPs with couplings large enough to solve the solar abundance problem truly existed, giving rise to observable effects in the frequencies of p-mode (sound wave) oscillations (Schlattl, Weiss & Raffelt 1999) and the solar neutrino flux (Gondolo & Raffelt 2009). Conversion in the Sun can also be directly probed: pointing an ‘axion helioscope’ at the Sun give strict limits on the reconversion of ALPs into X-rays (Arik et al. 2009). The CERN Axion Solar Telescope bound on Solar core ALP production and reconversion is gaγ < 9 × 10−11 GeV−1 for masses less than 10−2 eV.

  • Light shining through wall (LSW) experiments probe reconversion of ALPs produced from laser beams in an intense magnetic field. Current bounds are from the any light particle search (ALPS; Ehret et al. 2010) and GammeV (Chou et al. 2008) experiments. Both shine a high-intensity 532 nm laser beam at a stopper in a transverse magnetic field, and set bounds based on the (non) observation of reconverted photons behind the stopper. For masses below 1 meV, the ALPS bound is gaγ ≲ 6 × 10−8 GeV−1 whereas the GammeV bound is gaγ ≲ 3 × 10−7 GeV−1.

  • Vacuum polarization changes from ALP conversion of a linearly polarized infrared laser beam in a 2.3 T magnetic field has been probed by the polarizzazione del vuoto con laser (PVLAS) experiment (Bregant et al. 2008). Vacuum dichroism occurs when some of the photons oscillate into ALPs, causing a loss of part of the beam with polarization parallel to the magnetic field, and thus cause a small beam rotation. Conversely, birefringence occurs when ALP oscillation causes a phase lag of one of the polarization components, causing an ellipticity to develop in the beam polarization. The strongest bound to date from this effect has been provided by the dichroism limits of PVLAS: gaγ ≲ 4 × 10−7 GeV−1 for ALP masses below ∼1 meV.

These limits are much too strong to allow standard axions or ALPs to produce an observable effect in the solar atmosphere. Nonetheless, there are classes of models in which some of these bounds do not apply. It was pointed out by Masso & Redondo (2005) that if the ALP is a composite object analogous to the neutral pion, then the photon-ALP conversion can be suppressed at high energies when one of the photons is virtual. This suppresses ALP production in the hot interiors of stars, alleviating constraint 1. If the ALPs are furthermore short-lived particles, for instance if they decay to a lighter particle that does not interact strongly with the SM, then constraint 2 may also be overcome. In the latter case, the reconversion discussion in Section 3 becomes irrelevant.

However, dichroism constraints from PVLAS cannot be brushed aside even for non-SM of ALPs that can avoid constraints 1 and 2. Indeed, the photon-loss that would be observed in a dichroism experiment is the same as the effect that would be responsible for photon loss in the line-forming region of the solar atmosphere. If ALPs must respect the PVLAS upper limit on their coupling (gaγ ≤ 10−7 GeV−1), then the coupling required to produce an observable effect is suppressed by at least eight orders of magnitudes. We therefore conclude that current limits rule out axions and ALPs as progenitors of any observable effects on abundances derived from the solar photosphere.

Chameleons and hidden photons

In addition to ALPs, there exist further models of particles which mix with the photon in an analogous manner to (3), and one could hope that such models may allow constraints to be evaded. We look at two specific models: chameleons and hidden photons.

The chameleon, first proposed as a model of dark energy (Khoury & Weltman 2004a,b), consists of a scalar field ϕ with an effective coupling to matter |$V(\phi ) \sim {\rm e}^{\beta \phi /M_{\mathrm{Pl}}} \rho$|⁠, where β is a parameter of the theory, MPl is the Planck mass and ρ is the matter density. This gives rise to a particle (Davis, Schelpe & Shaw 2009; Brax & Zioutas 2010; Brax, Lindner & Zioutas 2012) whose mass depends on the local density via |$m_\phi ^2 \propto \rho$|⁠. Expanding the exponential coupling, the chameleon's first-order coupling to the photon is
\begin{equation} \mathcal {L}_{\phi \gamma } = \frac{\phi }{M_{\phi \gamma }} F_{\mu \nu }F^{\mu \nu }, \end{equation}
(28)
where Mϕγ = MPl/β is the mass scale of new physics giving rise to the chameleon field. The only relevant phenomenological difference with ALPs will be that chameleons could couple to photon polarization which is orthogonal, rather than parallel to, the bulk magnetic field, although nothing precludes a pseudo-scalar coupling. As we assume random polarizations here, this does not affect conversion in the solar atmosphere. The most recent constraint on the photon–chameleon coupling gives 1/Mϕγ < 3 × 10−8 GeV−1 (Steffen et al. 2010).
The hidden photon (or paraphoton) paradigm (Redondo 2008; Arkani-Hamed et al. 2009; Bjorken et al. 2009; Abrahamyan 2011; Essig et al. 2011) has its origins in an additional U(1) gauge field, with field strength F which mixes very weakly with the SM photon via a term
\begin{equation} \mathcal {L}_{\gamma ^{\prime }\gamma } = \frac{1}{2}\chi F_{\mu \nu }^{\prime } F^{\mu \nu }. \end{equation}
(29)
A combination of constraints from the cosmic microwave background, Coulomb-interaction, ‘LSW’ experiments and helioscopic observations constrain the mixing parameter χ ≲ 10−6 (Jaeckel, Redondo & Ringwald 2008; Mirizzi, Redondo & Sigl 2009; Ehret et al. 2010).
Crucially, the form of the coupling to photons means that resonant conversion of photons to either chameleons or hidden photons occurs in the same qualitative way as conversion to ALPs:
\begin{equation} P_{\gamma \rightarrow \phi _i} \propto \frac{1}{\left(m_{\phi _i}^2 - m_{\gamma ,\mathrm{eff}} ^2\right)^2 + \mathrm{const.}}. \end{equation}
(30)
Resonant oscillation is then forbidden for either of these scenarios at optical depths log τ ≲ 0, due to the dominance of |$\Delta _{||}^{\mathrm{gas}}$|⁠. Below such atmospheric heights, we find that the allowed parameter space of either theory does not permit oscillation. The hidden photon mixing parameter is roughly equivalent to |$\chi \leftrightarrow g_{a \gamma }B \omega /m_a^2$|⁠, whereas the chameleon coupling can be read directly as 1/Mϕγgaγ. Required parameters in both cases fall far above the current exclusion bounds, guaranteeing that no observable conversion may take place in the solar atmosphere.4

CONCLUSIONS

We have shown that for appropriate masses and couplings, particles with an axion-like interaction (equation 3) could in principle resolve the solar abundance problem, by inducing a slight increase in the effective continuum opacity at line-forming heights in the solar atmosphere. This in turn would reduce the computed equivalent widths of solar absorption lines for any given elemental abundance, and could bring inferred photospheric abundances into agreement with helioseismological results.

However, we found that the coupling necessary to obtain such a drastic change is ruled out by current experimental null results. We have shown that chameleons and hidden photons, which exhibit a similar phenomenological behaviour to the ALP, are also disfavoured as a photospheric solution to the solar abundance problem. We therefore conclude that new light bosons are not able to provide a photospheric solution to the solar abundance problem.

It is a pleasure to thank Martin Asplund, Remo Collett, Keshav Dasgupta, Anne Davis, Guy Moore Javier Redondo and Konstantin Zioutas for helpful comments and conversations. ACV was supported by NSERC, FQRNT and European contracts FP7-PEOPLE-2011-ITN, PITN-GA-2011-289442-INVISIBLES. PS was supported by the Lorne Trottier Chair in Astrophysics, an Institute for Particle Physics Theory Fellowship and a Canadian Government Tri-Agency Banting Fellowship, administered by NSERC. RT was supported by NASA grants NNX08AI57G and NNX11AJ36G.

1

A more complicated but related scenario was considered by Zioutas, Semertzidis & Papaevangelou (2007), where high-energy axions produced in the solar core would reconvert to photons in the photosphere and exert radiation pressure preferentially on heavier elements, driving them from the photosphere.

2

More correctly, the scaleheight of the line-to-continuous opacity ratio ηλ. As this scaleheight varies on a line-by-line basis, it is not especially useful for estimating effects on abundances as a function of formation height, in the transition-independent manner we do here.

3

The notation Maγ ≡ 1/gaγ is often used in such discussions.

4

This is true for the most optimistic case in which the density-dependent mass of the chameleon field allows resonance around log τ = 0. If this is not the case, resonance will of course not occur at all.

REFERENCES

Abrahamyan
S.
, et al. 
Phys. Rev. Lett.
2011
, vol. 
107
 pg. 
191804
 
Allende Prieto
C.
Lambert
D. L.
Asplund
M.
ApJ
2001
, vol. 
556
 pg. 
L63
 
Allende Prieto
C.
Lambert
D. L.
Asplund
M.
ApJ
2002
, vol. 
573
 pg. 
L137
 
Arik
E.
, et al. 
J. Cosmol. Astropart. Phys.
2009
, vol. 
0902
 pg. 
008
 
Arkani-Hamed
N.
Finkbeiner
D. P.
Slatyer
T. R.
Weiner
N.
Phys. Rev. D
2009
, vol. 
79
 pg. 
015014
 
Arnett
D.
Meakin
C.
Young
P. A.
Barnes
T. G.
III
Bash
F. N.
ASP Conf. Ser. Vol. 336, Cosmic Abundances as Records of Stellar Evolution and Nucleosynthesis
2005
San Francisco
Astron. Soc. Pac.
pg. 
235
 
Asplund
M.
ARA&A
2005
, vol. 
43
 pg. 
481
 
Asplund
M.
Deng
L.
Chan
K. L.
Proc. IAU Symp. 252, The Art of Modeling Stars in the 21st Century. Cambridge Univ. Press, Cambridge
2008
pg. 
13
 
Asplund
M.
Grevesse
N.
Sauval
A. J.
Allende Prieto
C.
Kiselman
D.
A&A
2004
, vol. 
417
 pg. 
751
 
Asplund
M.
Grevesse
N.
Sauval
A. J.
Barnes
T. G.
III
Bash
F. N.
ASP Conf. Ser. Vol. 336, Cosmic Abundances as Records of Stellar Evolution and Nucleosynthesis
2005a
San Francisco
Astron. Soc. Pac.
pg. 
25
 
Asplund
M.
Grevesse
N.
Sauval
A. J.
Allende Prieto
C.
Blomme
R.
A&A
2005b
, vol. 
431
 pg. 
693
 
Asplund
M.
Grevesse
N.
Sauval
A. J.
Scott
P.
ARA&A
2009
, vol. 
47
 pg. 
481
 
Asztalos
S. J.
Rosenberg
L. J.
van Bibber
K.
Sikivie
P.
Zioutas
K.
Ann. Rev. Nucl. Part. Sci.
2006
, vol. 
56
 pg. 
293
 
Ayres
T. R.
ApJ
2008
, vol. 
686
 pg. 
731
 
Ayres
T. R.
Plymate
C.
Keller
C. U.
ApJS
2006
, vol. 
165
 pg. 
618
 
Badnell
N. R.
Bautista
M. A.
Butler
K.
Delahaye
F.
Mendoza
C.
Palmeri
P.
Zeippen
C. J.
Seaton
M. J.
MNRAS
2005
, vol. 
360
 pg. 
458
 
Bahcall
J. N.
Basu
S.
Pinsonneault
M.
Serenelli
A. M.
ApJ
2005a
, vol. 
618
 pg. 
1049
 
Bahcall
J. N.
Serenelli
A. M.
Basu
S.
ApJ
2005b
, vol. 
621
 pg. 
L85
 
Bahcall
J. N.
Serenelli
A. M.
Basu
S.
ApJS
2006
, vol. 
165
 pg. 
400
 
Bartlett
M. L.
Power
E. A.
J. Phys. A
1969
, vol. 
2
 pg. 
419
 
Basu
S.
Antia
H.
ApJ
2004
, vol. 
606
 pg. 
L85
 
Basu
S.
Antia
H. M.
Phys. Rep.
2008
, vol. 
457
 pg. 
217
 
Bjorken
J. D.
Essig
R.
Schuster
P.
Toro
N.
Phys. Rev. D
2009
, vol. 
80
 pg. 
075018
 
Brax
P.
Zioutas
K.
Phys. Rev. D
2010
, vol. 
82
 pg. 
043007
 
Brax
P.
Lindner
A.
Zioutas
K.
Phys. Rev. D
2012
, vol. 
85
 pg. 
043014
 
Bregant
M.
, et al. 
Phys. Rev. D
2008
, vol. 
78
 pg. 
032006
 
Caffau
E.
Ludwig
H.-G.
Steffen
M.
Ayres
T. R.
Bonifacio
P.
Cayrel
R.
Freytag
B.
Plez
B.
A&A
2008
, vol. 
488
 pg. 
1031
 
Caffau
E.
Ludwig
H.-G.
Bonifacio
P.
Faraggiana
R.
Steffen
M.
Freytag
B.
Kamp
I.
Ayres
T. R.
A&A
2010
, vol. 
514
 pg. 
A92
 
Castro
M.
Vauclair
S.
Richard
O.
A&A
2007
, vol. 
463
 pg. 
755
 
Centeno
R.
Socas-Navarro
H.
ApJ
2008
, vol. 
682
 pg. 
L61
 
Charbonnel
C.
Talon
S.
Sci
2005
, vol. 
309
 pg. 
2189
 
Cheung
M. C.
Cameron
R. H.
ApJ.
2012
, vol. 
750
 pg. 
6
 
Chou
A. S.
, et al. 
Phys. Rev. Lett.
2008
, vol. 
100
 pg. 
080402
 
Christensen-Dalsgaard
J.
di Mauro
M. P.
Houdek
G.
Pijpers
F.
A&A
2009
, vol. 
494
 pg. 
205
 
Cumberbatch
D. T.
Guzik
J. A.
Silk
J.
Watson
L. S.
West
S. M.
Phys. Rev. D
2010
, vol. 
82
 pg. 
103503
 
Dasgupta
K.
Firouzjahi
H.
Gwyn
R.
J. High Energy Phys.
2008
, vol. 
0806
 pg. 
056
 
Davis
A.-C.
Schelpe
C. A. O.
Shaw
D. J.
Phys. Rev. D
2009
, vol. 
80
 pg. 
064016
 
Drake
J. J.
Testa
P.
Nat
2005
, vol. 
436
 pg. 
525
 
Duffy
L. D.
van Bibber
K.
New J. Phys.
2009
, vol. 
11
 pg. 
105008
 
Ehret
K.
, et al. 
Phys. Lett. B
2010
, vol. 
689
 pg. 
149
 
Essig
R.
Schuster
P.
Toro
N.
Wojtsekhowski
B.
J. High Energy Phys.
2011
, vol. 
2
 pg. 
9
 
Frandsen
M. T.
Sarkar
S.
Phys. Rev. Lett.
2010
, vol. 
105
 pg. 
011301
 
Gondolo
P.
Raffelt
G. G.
Phys. Rev. D
2009
, vol. 
79
 pg. 
107301
 
Grevesse
N.
Sauval
A. J.
Space Sci. Rev.
1998
, vol. 
85
 pg. 
161
 
Guzik
J. A.
Mussack
K.
ApJ
2010
, vol. 
713
 pg. 
1108
 
Guzik
J. A.
Watson
L. S.
Cox
A. N.
ApJ
2005
, vol. 
627
 pg. 
1049
 
Hayek
W.
Asplund
M.
Carlsson
M.
Trampedach
R.
Collet
R.
Gudiksen
B. V.
Hansteen
V. H.
Leenaarts
J.
A&A
2010
, vol. 
517
 pg. 
A49
 
Jaeckel
J.
Redondo
J.
Ringwald
A.
Phys. Rev. Lett.
2008
, vol. 
101
 pg. 
131801
 
Khoury
J.
Weltman
A.
Phys. Rev. D
2004a
, vol. 
69
 pg. 
044026
 
Khoury
J.
Weltman
A.
Phys. Rev. Lett.
2004b
, vol. 
93
 pg. 
171104
 
Masso
E.
Redondo
J.
J. Cosmol. Astropart. Phys.
2005
, vol. 
0509
 pg. 
015
 
Meléndez
J.
Asplund
M.
A&A
2008
, vol. 
490
 pg. 
817
 
Mirizzi
A.
Raffelt
G. G.
Serpico
P. D.
Kuster
M.
Raffelt
G.
Beltrá
B.
Lecture Notes in Physics, Vol. 741, Axions.
2008
Berlin
Springer-Verlag
pg. 
115
 
Mirizzi
A.
Redondo
J.
Sigl
G.
J. Cosmol. Astropart. Phys.
2009
, vol. 
0903
 pg. 
026
 
Moll
R.
Cameron
R.
Schussler
M.
A&A
2012
, vol. 
541
 pg. 
A68
 
Morel
T.
Butler
K.
A&A
2008
, vol. 
487
 pg. 
307
 
Peccei
R. D.
Quinn
H. R.
Phys. Rev. Lett.
1977
, vol. 
38
 pg. 
1440
 
Raffelt
G. G.
Phys. Rev. D
1988
, vol. 
37
 pg. 
1356
 
Raffelt
G. G.
Phys. Rep.
1990
, vol. 
198
 pg. 
1
 
Raffelt
G. G.
Ann. Rev. Nucl. Part. Sci.
1999
, vol. 
49
 pg. 
163
 
Raffelt
G. G.
Kuster
M.
Raffelt
G.
Beltrán
B.
Lecture Notes in Physics, Vol. 741, Axions.
2008
Berlin
Springer-Verlag
pg. 
51
 
Raffelt
G.
Stodolsky
L.
Phys. Rev. D
1988
, vol. 
37
 pg. 
1237
 
Redondo
J.
J. Cosmol. Astropart. Phys.
2008
, vol. 
7
 pg. 
8
 
Rutten
R. J.
Radiative Transfer in Stellar Atmospheres
2003
8th edn
Lecture Notes. Utrecht University
 
Schlattl
H.
Weiss
A.
Raffelt
G.
Astropart. Phys.
1999
, vol. 
10
 pg. 
353
 
Scott
P.
Asplund
M.
Grevesse
N.
Sauval
A. J.
A&A
2006
, vol. 
456
 pg. 
675
 
Scott
P.
Asplund
M.
Grevesse
N.
Sauval
A. J.
ApJ
2009
, vol. 
691
 pg. 
L119
 
Serenelli
A.
Basu
S.
Ferguson
J. W.
Asplund
M.
ApJ
2009
, vol. 
705
 pg. 
L123
 
Serenelli
A. M.
Haxton
W. C.
Peña-Garay
C.
ApJ
2011
, vol. 
743
 pg. 
24
 
Socas-Navarro
H.
Norton
A. A.
ApJ
2007
, vol. 
660
 pg. 
L153
 
Steffen
J. H.
Upadhye
A.
Baumbaugh
A.
Chou
A. S.
Mazur
P. O.
Tomlin
R.
Weltman
A.
Wester
W.
Phys. Rev. Lett.
2010
, vol. 
105
 pg. 
261803
 
Stein
R. F.
Nordlund
Å.
ApJ
1998
, vol. 
499
 pg. 
914
 
Stein
R. F.
Nordlund
A.
ApJ
2006
, vol. 
642
 pg. 
1246
 
Steiner
O.
Rezaei
R.
Schlichenmaier
R.
Schaffenberger
W.
Wedemeyer-Böhm
S.
Lites
B.
Cheung
M.
Magara
T.
Mariska
J.
Reeves
K.
ASP Conf. Ser. Vol. 415, The Horizontal Magnetic Field of the Quiet Sun: Numerical Simulations in Comparison to Observations with Hinode
2009
San Francisco
Astron. Soc. Pac.
pg. 
67
 
Svrcek
P.
Witten
E.
J. High Energy Phys.
2006
, vol. 
0606
 pg. 
051
 
Taoso
M.
Iocco
F.
Meynet
G.
Bertone
G.
Eggenberger
P.
Phys. Rev. D
2010
, vol. 
82
 pg. 
083509
 
Van Bibber
K.
Dagdeviren
N. R.
Koonin
S. E.
Kerman
A. K.
Nelson
H. N.
Phys. Rev. Lett.
1987
, vol. 
59
 pg. 
759
 
Yang
W. M.
Bi
S. L.
ApJ
2007
, vol. 
658
 pg. 
L67
 
Zioutas
K.
Semertzidis
Y.
Papaevangelou
T.
Axebides
M.
Fanourakis
G.
Vergados
J.
The Identification of Dark Matter. World Scientific Publishing, Singapore
2007
pg. 
372