CONDENSED MATTER ASTROPHYSICS: A PRESCRIPTION FOR DETERMINING THE SPECIES-SPECIFIC COMPOSITION AND QUANTITY OF INTERSTELLAR DUST USING X-RAYS

, , , , and

Published 2009 August 18 © 2009. The American Astronomical Society. All rights reserved.
, , Citation Julia C. Lee et al 2009 ApJ 702 970 DOI 10.1088/0004-637X/702/2/970

0004-637X/702/2/970

ABSTRACT

We present a new technique for determining the quantity and composition of dust in astrophysical environments using <6 keV X-rays. We argue that high-resolution X-ray spectra as enabled by the Chandra and XMM-Newton gratings should be considered a powerful and viable new resource for delving into a relatively unexplored regime for directly determining dust properties: composition, quantity, and distribution. We present initial cross section measurements of astrophysically likely iron-based dust candidates taken at the Lawrence Berkeley National Laboratory Advanced Light Source synchrotron beamline, as an illustrative tool for the formulation of our technique for determining the quantity and composition of interstellar dust with X-rays. (Cross sections for the materials presented here will be made available for astrophysical modeling in the near future.) Focused at the 700 eV Fe Liii and Lii photoelectric edges, we discuss a technique for modeling dust properties in the soft X-rays using L-edge data to complement K-edge X-ray absorption fine structure analysis techniques discussed by Lee & Ravel. The paper is intended to be a techniques paper of interest and useful to both condensed matter experimentalists and astrophysicists. For the experimentalists, we offer a new prescription for normalizing relatively low signal-to-noise ratio L-edge cross section measurements. For astrophysics interests, we discuss the use of X-ray absorption spectra for determining dust composition in cold and ionized astrophysical environments and a new method for determining species-specific gas and dust ratios. Possible astrophysical applications of interest, including relevance to Sagittarius A*, are offered. Prospects for improving on this work in future X-ray missions with higher throughput and spectral resolution are also presented in the context of spectral resolution goals for gratings and calorimeters, for proposed and planned missions such as Astro-H and the International X-ray Observatory.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Understanding dust is vital to our understanding of the Universe. Dust is a primary repository of the interstellar medium (ISM) and contributes to the chemical evolution of stars, planets, and life itself. A better understanding of the content of astrophysical dust has far reaching relevance and application to areas of astrophysics ranging from nucleosynthesis to planet formation. Dust is everywhere, and for many astrophysical topics, its presence hinders our studies in some way, be it to extinguish a UV part of a spectrum, or possibly affect Cosmology results, as, e.g., in the case of using Type Ia supernovae light curves (which are affected by line-of-sight dust) as a beacon for probing the dark energy content in the Universe. Therefore, at a minimum, a better understanding of dust will allow us to better isolate its effects (be it complicating spectra or light curves) to get at a cleaner study of topics from cosmology to black hole environments. There is a multiwavelength industry (mostly radio to IR) focused on the problem of astrophysical dust, and in recent years, Spitzer observations have significantly improved our understanding. Yet, despite good progress, there remains much to understand about dust properties (size, composition, and distribution) in astrophysical environments, from the colder ISM environs to the hotter environments near the disks and envelopes of young stars and compact sources. Our quest to address outstanding problems, therefore, can greatly benefit from additional complementary and/or orthogonal techniques.

There are several advantages of studying dust properties in the X-rays, the most significant of which is that we can directly measure its quantity (relative to gas phase), and composition. ("Dust" in this paper includes complex molecules). In the UV and optical wavelength, the presence of dust is generally inferred through depletion, rather than directly measured. Radio spectroscopy is able to probe the presence of certain molecules, and IR, some subset of ISM compounds (mostly polycyclic aromatic hydrocarbon or PAHs, graphites, certain silicates, and ice mantle bands). In these bands, spectral features come from probing the molecule as a whole through rotational or vibrational modes originating from, e.g., the excitation of phonons (rather than electrons). In contrast, both gas and (≲10 μm) dust are semitransparent to X-rays, so the measured absorption in this energy band is sensitive to all atoms in both gas and solid phase. Therefore, these high-energy photons can be used to facilitate a direct measurement of condensed phase chemistry via a study of element-specific atomic processes, whereby the excitation of an electron to a higher lying quantum level, band resonance, or molecular structure will imprint modulations on X-ray spectra that reflect the individual atoms that make up the molecule or solid. Therefore, in the same way that we can look at a given energy to identify ions, the observed spectral modulations near photoelectric edges, known as X-ray absorption fine structure (XAFS), provide unique signatures of the condensed matter that imprinted that signature. For this reason, high-resolution X-ray studies, currently best enabled by X-ray grating observations (Chandra and XMM), provide a unique and powerful tool for determining the state and composition of ISM grains.

Lee & Ravel (2005) have discussed the viability of XAFS analysis and techniques for K-edge data with particular emphasis on iron in the hard >7 keV X-rays, in early anticipation of the launch of the Suzaku calorimeters. (See also Woo 1995, Woo et al. 1997, and Forrey et al. 1998 for early theory discussions relating to XAFS detections in astrophysical environments.) Due to the unfortunate in-flight failure of the instrument, the viability of astrophysical XAFS studies will have to focus on the softer <6 keV regime with existing instruments, until the Astro-H launch. While a more complicated part of the spectrum (due to the numerous ions that populate this spectral region), XAFS signatures in the soft X-rays have been reported in early papers based on Chandra and XMM spectral studies (Paerels et al. 2001; Lee et al. 2001, 2002; Ueda et al. 2005; Kaastra et al. 2009; de Vries & Costantini 2009; see also Takei et al. 2002; Schulz et al. 2002; Juett et al. 2004, 2006 for studies focused on the oxygen K edge for abundance determinations). It is the detection of these features in X-ray spectra of astrophysical objects that have motivated us to obtain laboratory measurements of likely ISM grain candidates to facilitate astrophysical modeling efforts. To date, no modeling of the magnitude we propose here has been undertaken. The ∼0.1–8 keV energy coverage combined between the Chandra and XMM-Newton gratings will allow us to look in more detail at photoelectric edges near C K, O K, Fe L, Mg K, Si K, Al K, S K, Ca K, and Fe K, and therefore molecules/grains containing these constituents. We note, however, that for the K-edges of C, S, Ca and Fe, we are likely to be limited by signal-to-noise ratio (S/N) and/or spectral resolution, but are nevertheless well situated to study magnesium-, silicate, and aluminum-based grains using K-edge data, and iron-based grains/molecules can be studied using L-edge spectra at ∼700 eV. Additionally, because X-ray absorption spectra are of some admixture of gas and dust, each having very different individual spectral signatures, we can separate their relative contributions through spectral modeling. Our proposed technique for doing so is discussed in Section 3, using cross section measurements of XAFS at Fe L as the illustrative example for our discussion points.

This paper presents dialog focused on the application of condensed matter X-ray techniques to the soft X-ray absorption study of astrophysical dust properties: composition and quantity. What knowledge we gain from this work will greatly complement current IR studies of dust. Ultimately, the questions we would strive to address in this study include: (1) the mineralogy of dust in different astrophysical environments, and perhaps (2) something of its nature (e.g., crystalline or amorphous). Since K-edge analysis has already been discussed by Lee & Ravel (2005), we focus on XAFS studies as applied to L-edges, where the spectral modeling of near-edge absorption modulations (XANES: X-ray absorption near-edge structure), rather than far-edge absorption modulations (EXAFS: extended X-ray absorption fine structure) play a primary role. While the emphasis is on the ∼700 eV Fe L spectral region, the techniques presented will be widely applicable to all L-edge studies. Because condensed matter measurements are discussed in the context of astrophysics studies, we organize this paper into sections that would be of interest and useful to both condensed matter experimentalists (Section 2) and astrophysicists (Sections 3 and 4). For the experimentalists, Section 2.2 may be of particular interest for its discussion of an L-edge cross section normalization technique that is robust to laboratory synchrotron cross section measurements with relatively poor S/N. For astrophysics interests, Section 3 discusses a new method for determining interstellar dust quantity and composition through X-ray studies. This is followed in Section 4 with a brief discussion of some possible studies of interest and ability of this technique to separate out fractional contributions of different composition of dust in different regions. Prospects for improving on this work in the context of future missions are discussed in Section 5.

2. THE SAMPLES AND LABORATORY EXPERIMENT

UV, IR, and planetary studies of meteorites and dust have pointed to interstellar dust grains condensed from heavier elements such as C, N, O, Mg, Si, and Fe. With respect to using Chandra High Energy Transmission Grating Spectrometer (HETGS) spectra to enable detailed studies of absorption structure near photoelectric edges to determine grain composition, Fe-, Mg-, and Si-based dust would be the most relevant since the L-edge of Fe and K-edges of Mg and Si are all contained within the ∼0.5–8 keV (∼1.5–25 Å) HETGS bandpass, where the effective area is also highest. (The Fe K-edge is also encompassed by the HETGS, but the S/N there is usually too low to be useful for analysis.) For this paper, we compare cross section measurements for hematite (α-Fe2O3), iron sulfate (FeSO4), fayalite (Fe2SiO4), and lepidocrocite (γ-FeOOH). Cross section measurements at Fe L and Si K are nearing completion for the most astrophysically likely iron- and silicon-based dust, which will be presented in a future paper on XAFS standards for astrophysical use, while the data presented in this paper are used primarily to illustrate our proposed analysis techniques and methodology. Planned measurements are also being made for other interesting edges discussed previously.

The laboratory data presented in this paper were taken at the Lawrence Berkeley National Laboratory Advanced Light Source beamline 6.3.1. This is a bending magnet beamline with a Variable Line Spacing Plane Grating Monochromator (VLS-PGM) that is sensitive to 300–2000 eV energy range and has a resolving power E/ΔE = 3000. This resolution exceeds the E/ΔE = 1000 Chandra HETGS resolving power by factor of three.

The 6.3.1 detectors allow for a choice between total electron yield, photodiode (fluorescence), and transmission experiments. The latter would have been the ideal choice, since transmission measurements would be directly related to what we seek to understand, namely the transmission of X-rays through interstellar grains of comparable thickness. However, a nominal thickness estimate for 30% transmission would require film/foil thicknesses for our candidate samples in the range t∼ 0.1–0.8 μm from the transparent to opaque side of the edge at Fe L, as calculated according to the formalism T = e−μρt. (The density ρ for the compounds was taken from "The Handbook of Chemistry and Physics," and the values for the attenuation length were obtained from the6 CXRO at LBL and7 NIST.) Preparation of pure-phase samples of that thickness was impractical. As such, we opted to obtain cross section measures using both fluorescence and electron yield (Section 2.1).

2.1. Fluorescence and Electron Yield Experiments

In an X-ray-absorption spectroscopy (XAS) experiment, the incident X-ray photon promotes a deep-core electron into an unoccupied state above the Fermi energy. For the iron L edge experiments shown in this paper, that deep core state is a 2p state. As the incident photon energy is varied using a variable line spacing plane grating monochromator, the cross section of this absorption process is monitored by measurement of the decay of a higher-lying electron into the core hole. The fluorescence of a photon and emission of Auger and secondary electrons can be monitored in tandem. Because the mean-free path of the electron is limited to a few nanometers or less, the electron yield measurement is mostly sensitive to the surface of the sample. The fluorescent photon, on the other hand, has a larger mean-free path, and therefore is a better probe of the bulk of the sample.

The samples measured in this paper were fine powders, either purchased as such or prepared by grinding in an agate mortar and pestle. A portion of these fine powders was sprinkled onto conducting carbon tape, another portion was pressed onto a flattened strip of indium metal. The carbon tape and indium strips were affixed to a glass microscope slide for mounting onto the sample holder. The fluorescence measurement is made using a photodiode pointed at the sample. However, since all of our samples were thick compared to the penetration depth and highly concentrated in iron, the fluorescence data were significantly attenuated by the self-absorption effect (see, e.g., Booth & Bridges 2005). All data in this paper were measured in electron yield mode.

To measure the electron yield, an alligator clip was clamped onto the glass slide and electrical contact was made with the otherwise isolated carbon tape and indium strip, and used to carry the photocurrent into a current-to-voltage amplifier. This experimental arrangement provided for redundant measurements of each sample. Some samples proved more amenable to measurement on the carbon tape substrate and others to the indium. For every sample, each measurement was repeated three times. In each case, the measurement yielding the highest quality, most reproducible data was selected for presentation. The three scans for each sample have been calibrated in energy, aligned, and averaged.

Absolute energy calibration was attained by measuring an iron foil and calibrating it to the tabulated energies of the iron LII and LIII edges (Brennan & Cowen 1992). We had no capacity to prepare the surface of our foil, thus we relied upon the fluorescence measurement. Although the foil measurement is severely attenuated by self-absorption, the edge position can be reliably extracted. Iron-edge scans on other materials were aligned to the iron foil data using reproducible features in the spectrum of the incident intensity.

2.2. Absorption Cross Section Normalization Technique

In an XAS measurement, the absolute scale of the measured cross section is ambiguous. The size of the measured step at an absorption edge energy depends on a wide variety of factors, including the concentration of the absorbing atom in the sample, the electronic gains on the signal chains, and the efficiencies of the detectors. In order to compare measurements on disparate samples under differing experimental conditions, it is standard practice to perform a common normalization of every measured spectrum. In this work, we choose to perform this normalization by matching the measurements to tabulated values for bare-atom cross sections, as described in this section. In this way, we can directly and consistently compare measurements on our samples to one another as well as to astronomical observations.

The normalization of XAS L edge data is more complicated than for K edge data due to the close proximity of the LII and LIII edges in the soft X-ray regime. In order to properly normalize these spectra, we adopt the MBACK normalization technique of Weng et al. (2005) with modifications. Here, we describe the MBACK technique in brief, and our modification, which we found to be necessary for data with ≳0.2% noise, as typical of each individual cross section measure. We note that the typical practice has been to first co-add all individual cross section measures to increase S/N, before normalization. However, we wished to be more rigorous in our methodology by first renormalizing the individually measured cross sections before averaging to create the final cross section to facilitate fitting purposes. The MBACK technique ensures better normalization for the more complex L-edge cross sections by calculating a smooth normalizing function over the entire measured data range, rather than extrapolating from independently pre- and post-edge linear fits as would be sufficient for the less complex K-edge cross sections. For our data, this normalizing function, μnorm, is dominated by the Legendre polynomial term (in brackets) of the function

Equation (1)

where Eedge is the energy of the onset of the edge absorption, and m and Cj are, respectively, the polynomial order and coefficient. Weng et al. additionally note that a rapidly decreasing pre-edge shape, due to residual elastic scattering, is often seen in fluorescence experiments using energy discriminating detectors, and therefore include an additional error function component in the normalization as defined by the nonbracketed term of Equation (1); see Figure 1 for the shape of this function as determined from our sample fit to γ-FeO(OH)−lepidocrocite. Here, Eem and ξ are, respectively, the centroid energy of the X-ray emission line of the absorbing element and its width, with A as a scale factor. We note, however, that this function is adopted here solely as an effective way to model similar pre-edge feature shapes originating from experimental artifacts of uncertain origin. (Residual elastic scattering effects do not apply to our experiment, which measured total photocurrent.) In any case, our data do not show a large rapidly decreasing pre-edge slope, so this term has negligible effect on our overall normalization; nevertheless, we include it in our modeling efforts for completeness, since this would allow for a global prescription for the determination of μnorm that is independent of experimental technique. See Figure 2(a) for an illustration of the components that make up the normalizing function for our sample γ-FeO(OH) compound. In general, the details of the normalization variables would clearly depend on the composition of the condensed matter.

Figure 1.

Figure 1. Error function as defined in Equation (1), which accounts for any decreasing function in the pre-edge slope due to residual inelastic scattering.

Standard image High-resolution image
Figure 2.

Figure 2. Illustration based on γ-FeO(OH) of the steps involved in the normalization of Fe L edge synchrotron data as detailed in Section 2.2 and Equations (1) and (3). (a) Our measured cross section scaled by s (black), plotted with the best determined normalization μnorm (magenta), as defined in Equation (1) to be the sum of μpoly (the dashed dark blue) and μerfc (the dashed-dot light blue). (b) The pre- and post-edge regions are then normalized against the tabulated cross sections μCL of Brennan & Cowen (1992), which are shown here as a step function arbitrarily enhanced to five times its value (the dash-dot green), for illustrative purposes. Shown in orange solid is the μCLback component of the Equation (3) minimizing function. (c) The normalized γ-FeO(OH) cross sections (red) representing bound–bound absorption, plotted with the μCL absorption cross sections (green-dashed) representing bound-free continuum absorption by Fe L; also plotted is the prenormalization data of panel (a) (black).

Standard image High-resolution image

To determine the best normalization (Figure 2(b)), we initially employ the prescribed minimizing function of Weng et al.

Equation (2)

where N represents the total number of data bins, i is the bin index, and n1 and n2 are, respectively, the total number of pre-edge and post-edge data bins, such that Nn2 + 1 represents index 1 of the post-edge bin. The variable μbf is the Brennan & Cowen (1992) tabulated absorption cross sections representing the bound-free nonresonant absorption component of Fe. Note that μraw is our measured cross sections, which is scaled by s, and μnorm is the normalization function of Equation (1). However, in our fitting efforts, we found the scale factor s to be very sensitive to noise, often taking on small values for our data which resulted in badly normalized cross sections. In considering a remedy for this, we tested Equation (2) based on simulated cross sections (μbf + μnorm), where we have added controlled levels of Gaussian-distributed noise, as per the prescription (1 + f Gnoise) · (μbf + μnorm), where Gnoise is a group of random numbers following the Gaussian distribution and modified by the fractional level f, i.e., f = 0.2% or f = 1%, as in Figure 3. (The Gaussian distribution function is defined as $\phi (x)={e^{-(x-\mu)^2/2\sigma ^2}} / \sigma \sqrt{2\pi }$ where the expectation value μ = 0 and the standard deviation σ = 1.) (The μnorm parameters of Equation (1) are set to C0 = 3.0, C1 = 0.01, C2 = 0.0002, A = 1.0, Eem = 685 eV, ξ = 5.0, Eedge = 707 eV). We then divide the simulated data by the value of s arbitrarily set to 30, before fitting according to the minimization function of Equation (2) to see if we can recover the same value. In doing so, we find that the Weng et al. method, sans modification, is robust only for data with noise at ≲ 0.2%. Accordingly, we modify Equation (2) by dividing sμraw(Ei) into the pre-edge and post-edge minimization terms as

Equation (3)

and find that our method (Equation (3)) is robust to data of up to 2% noise. In total, the normalization procedure will use m + 5 adjustable parameters (m + 1 polynomial coefficients, two scale factors of s and A as well ξ and Eem) to determine the best normalization for the measured cross sections.

Figure 3.

Figure 3. Comparison of the normalization technique for compound γ-FeOOH, based on Equation (2) (normalized data based on the method of Weng et al. 2005 shown in blue; arbitrarily scaled up for illustrative purposes) versus its modification (Equation (3)) as proposed in this paper (normalized data shown in red) for simulated data (green) with varying levels of Gaussian-distributed noise. The Fe L cross sections of Brennan & Cowen (1992) are shown in black to illustrate goodness of fit. The normalization function of Equation (1) is shown in magenta.

Standard image High-resolution image

Of relevance to the cross sections presented in this paper, 18 measurements (nine mounted on indium substrate and nine on carbon) of each sample were taken. However, the carbon substrate measurements were consistently nosier, so we exclude those from the final co-averaged cross sections. For each individual indium substrate cross section measure (see Figure 4, top), the data were normalized using Equations (1) and (3) based on the subsequent steps. To compare with our reference spectrum, each measurement is first shifted by 4.1 eV to correct for the calibration offset of the monochromator. The resultant energy span of our measurements is then 684.1–754.1 eV, which corresponds to N = 700 total data bins of 0.1 eV widths. As described, Equations (1) and (3) were then used to determine the best normalized description for the combined data set unique to each sample. Since all data measured spanned the range noted above, the number of pre-edge and post-edge data bins correspond respectively to n1 = 179 (684.1–702 eV) and n2 = 141 (740–754.1 eV) for all samples. The edge energy Eedge = 707 eV corresponding to the Fe L iii, and a polynomial of order m = 2 was deemed sufficient, i.e., the goodness of fit showed negligible improvement for higher polynomial orders. Postnormalization, cross sections of the highest quality were co-added to create the final cross sections of Figure 5. For the final cross sections, no fewer than nine (indium substrate) data sets were co-averaged per compound type. Figure 5 shows our measured cross sections post-normalization, compared against metallic (Kortright & Kim 2000), and what we assume for bound-free continuum absorption by iron (Brennan & Cowen 1992; Henke et al. 1993). For future fitting efforts to astrophysical data, we use the structure-less Fe L tabulated values of Brennan & Cowen (1992) to extrapolate the normalized XAFS data beyond the energy span of our measurements to encompass an energy range that extends from 0.4 keV to 10 keV.

Figure 4.

Figure 4. Comparison of the cross section normalization technique proposed in this paper (top; Equation (3)) versus that proposed by Weng et al. 2005 (Equation (2); bottom) for the sample compound γ-FeO(OH). In each panel, individual cross section measurements (arbitrarily scaled higher to facilitate plotting purposes) are shown in multicolors, while just below the combined normalized data according to the respective methods are shown in red. The blue bracketing is the standard deviation based on the individual measurements to give sense of measurement differences as a function of energy. The Brennan & Cowen (1992) cross sections are shown in black to compare how the normalization techniques differ.

Standard image High-resolution image
Figure 5.

Figure 5. X-ray absorption near edge structure in the vicinity of the Fe L photoelectric edge, postnormalization, reveals that structures known as XAFS are distinct for different states of condensed matter. Note also differences in edge structure between bound-free continuum absorption (the solid black step function) vs. metallic (the dashed black) and molecular (in color) states. Note that these are only preliminary measurements which we intend to improve upon before incorporation into astrophysical databases for common use. The inset illustrates the ΔE values distinguishing condensed matter (red) from gas-phase (black) absorption at the Fe LIII and LII photoelectric edge energies, for the different compounds which are tabulated in Table 1.

Standard image High-resolution image

We plan to avail the community of our cross section measurements8 for these and other compounds, as absolutely calibrated standards for interstellar dust studies, in the near future. Table 1 lists details for a few of these, but as stated, the data shown here are intended merely as a vehicle to facilitate a presentation of our new techniques.

Table 1. Photoelectric Edge Energies of Measured Compounds

Compoundsa Chargeb state EcLIII eV(Å) ΔEdLIII eV(Å) EeLII eV(Å) ΔEfLII eV(Å)
FeO(OH)-lepidocrocite Fe3+ 705.7(17.57) −2.7 (+0.07) 718.9(17.25) −2.3 (+0.06)
Fe2O3-hematite Fe3+ 705.7(17.57) −2.7 (+0.07) 718.9(17.25) −2.3 (+0.06)
Fe2SiO4-fayalite Fe2+ 704.3(17.60) −4.1 (+0.10) 717.4(17.28) −3.8 (+0.09)
FeSO4-iron sulfate Fe2+ 704.3(17.60) −4.1 (+0.10) 717.8(17.27) −3.4 (+0.08)

Notes. aLaboratory measurements of cross sections for astrophysically likely species presented in this paper. bThe charge state reflects the number of electrons removed from iron when the compound is formed. To take Fe2SiO4 (fayalite) as an example: silicon (atomic orbital: 1s22s22p63s23p2) will lose four electrons in its outmost shell so that oxygen (1s22s22p4) can fill its shell. Since there are four oxygen atoms, two additional electrons will be taken from Fe such that the charge state of the compound is +2. cThe Fe LIII photoelectric edge energy associated with the compound of interest, as measured at the 708.4 eV Fe LIII peak of the Brennan & Cowen (1992) tabulated cross sections. See Figure 5 inset for illustration. dThe energy difference between compound and bound-free absorption (as tabulated based on the cross sections of Brennan & Cowen 1992). eSame as column (c) but for the Fe LII photoelectric edge at 721.2 eV assumed for bound-free continuum absorption. fSame as column (d) but for the Fe LII photoelectric edge at 721.2 eV.

Download table as:  ASCIITypeset image

3. AN X-RAY METHOD TO DETERMINE THE QUANTITY AND COMPOSITION OF DUST IN INTERSTELLAR SPACE

Our ability to accurately measure the quantity of dust and elemental abundances in our Galaxy and beyond has far-reaching applications and consequences for a diverse range of astrophysical topics. Thus far, wavelength-dependent (IR to UV) studies of starlight attenuation as facilitated by E(BV) measurements and other extinction studies have been the primary technique by which we measure the amount of dust that is bound up in interstellar grains. Elemental depletion can be determined by UV absorption studies comparing the amount expected in gas-phase absorption from what is observed, and IR spectral studies can directly measure certain ISM dust. Here, we present an X-ray technique for directly determining the (element-specific) quantity and composition of interstellar gas (Section 3.1) and dust from within a single observation. What knowledge we gain from this technique, when combined with the wealth of knowledge from non-X-ray studies, will significantly increase our understanding of ISM dust and its effects on astrophysical environments.

3.1. Gas-phase ISM Absorption

Like condensed material, atomic transitions to higher quantum levels within an isolated atom will also give rise to multiplet resonant absorption. To give the example for Fe, these lines would consist of all possible discrete transitions to all possible configurations of the 3d-shell, since the electronic configuration of Z = 26 Fe+0 (or Fe i in the astronomy notation) is 1s22s22p63s23p64s23d6. Therefore, while the bound-free nonresonant transitions can be modeled by a simple step function as, for example, that of Brennan & Cowen (1992) for Fe L, additional discrete resonant features will have to be included to account for the bound–bound resonant component of absorption.

Figure 6 shows our calculations for the resonant transition for various Fe ions from Fe+0 − Fe+4, as evaluated based on the Gu et al. (2006; see also Gu 2005) predictions for oscillator strengths, radiative decay rates, and autoionization rates for these lines. We note that because of the electronic structure of Fe, which favors the removal of the s-electrons prior to d-electrons, the cross sections for Fe+0 ∼ Fe+1 ∼ Fe+2, as evident in the figure. The relevant cross sections derived for these atomic transitions are then convolved with the appropriate instrument resolution (ΔE ∼ 0.9 eV at 700 eV Fe L for the Chandra Medium Energy Grating) to approximate the ion-specific cross section for the bound–bound transition, σbb. Finally, the cross section for describing the total gas-phase absorption for the species of interest, σZgas, will be described by the combination of cross sections from the bound–bound (μbb − atom) and bound-free (μbf) components. See Section 3.2 for more details.

Figure 6.

Figure 6. Bound–bound resonant transition for various Fe ions from Fe+0 to Fe+4, as calculated based on the (Gu et al. 2006; Gu 2005) predictions for oscillator strengths, radiative decay rates, and autoionization rates for these lines. Based on the ISM ionization spectrum of Sternberg et al. (2002), the most prominent contribution from Fe ions in the L-edge region would come from neutral Fe+0 (i.e., Fe i), or singly ionized Fe+1 (Fe ii). These lines are convolved with the Chandra HETGS spectral resolution (R ∼ 0.9 eV at FeL; blue), and the resolution of ALS beamline 6.3.1 is used for the XAFS measurements presented in this paper (red). At present, R ∼ 3000 is also the baseline spectral resolution for the IXO spectrometers.

Standard image High-resolution image

For complex astrophysical environments, additional considerations related to the ionization state of the plasma being probed need to be weighed. As such, modeling efforts need to consider the ionization state of the environment expected for the absorption. For the three, cold (T < 200 K), warm (T ∼ 8000 K), and hot (T ∼ 5 × 105 K) phases of the ISM, the iron ions that contribute most strongly to the L-edge region are Fe+0 followed by Fe+1. At T = 8000 K, the peak ion fractions are Fe+0 ∼ 0.67, Fe+1 ∼ 0.33, Fe+2 ∼ 2.8 × 10−3 and Fe+3 ∼ 1.9 × 10−6, assuming the ISM ionizing spectrum defined by Sternberg et al. (2002). At the much colder temperatures of the cold ISM, only Fe+0 contributes, and at the much hotter temperatures of the hottest phase of the ISM, it is Fe+25 (at 6.7 keV) which contributes the bulk ion fraction.

3.2. A Prescription for Determining Species-specific Gas-to-Dust Ratios

As stated in Section 1, the cross sections we measure in the laboratory of astrophysically likely dust candidates (Section 2) will be applied to X-ray spectra showing structure near photoelectric edges to determine dust composition and quantity. We propose the following prescription to do so:

Equation (4)

whereby the absorption contributions from gas (cold and ionized) and dust can be separately accounted for by each of the exponential terms. Here, Ffinal and F0 are respectively the detected and incident flux, and T, represented by the exponential terms, is the transmission. The τionized-gas is determined through photoionization modeling to account for any additional lines arising from the plasma of the source, and should be treated on a source-by-source basis. However, in the Fe L iii and L ii edge regions, we note that additional line contributions would have to come from high-transition (low oscillator strength) lines of He-like O vii (O+6) 1s2 − 1snp for n ⩾ 4, and Fe i–Fe xi (i.e., Fe+0 − Fe+10). Any significant contribution from oxygen would have to come from very ionized optically thick plasma; for the iron, moderate temperature plasma would have to be present in large quantities to have any appreciable effect in this spectral region. Therefore, for many astrophysical situations, neither ionized oxygen nor iron from the source plasma is expected to contribute strongly (if at all) to the Fe L iii and Fe L ii spectral region. Nevertheless, as stated, these can be modeled out through photoionization studies, the discussion of which is beyond the scope of this paper. For the total line-of-sight (LOS) (gas + dust) ISM absorption from all heavy elements, excluding the species of interest,

Equation (5)

where $\sigma _{Z^{\prime }}$, Asolar, and NH are respectively defined to be the cross section, abundance, and line of sight equivalent Hydrogen column summed over all heavy elements present in the broadband X-ray spectrum, with the exception of the element we are trying to measure. One reason for removing the species of interest from the broadband fitting is that we are interested in decomposing the gas and dust contribution specific to that species. For this, we include two additional exponential terms to describe the absorption by the gas and dust for the species of interest whereby

Equation (6)

and

Equation (7)

In these equations, σZgas and σZdust refer to the cross section of the gas and dust, respectively, for the species of interest. The determination of σZgas has been discussed in Section 3.1. The σZdust and our efforts to measure this for astrophysically viable dust candidates have been discussed at length in Section 2. Given enough S/N, the species-specific abundance AZ can be fitted for, with the condition that the total abundance should be a combination of both gas and dust, hence the additional multiplicative factor to AZ of x and (1 − x). Figure 7 shows these components as separated out in a fit to the X-ray binary Cygnus X-1. Details about the dust in Cygnus X-1 will be separately discussed in J. C. Lee et al. (2009, in preparation).

Figure 7.

Figure 7. The 700 eV Fe L photoelectric edge region as observed in the X-ray binary Cygnus X-1 reveals how we can decompose the absorption components to determine the quantity and composition of dust. The best fit (red) shows the combined absorption from all gas (blue) which includes line-of-sight cold gas (magenta) and XAFS signifying condensed matter (CM). Here, the ionized gas component contribution comes largely from plasma that is optically thick to He-like O vii, and therefore the 1s2 − 1snp (for n > 3) O vii resonance lines (blue) contribute to the absorption region co-located with the Fe Liii and L ii edges. In green is the absorption component representing the bound-free transition from the Fe ion and solid/molecule.

Standard image High-resolution image

4. DISCUSSION

The Chandra and XMM-Newton archives are rich with high-resolution spectra of X-ray bright binaries and active galactic nuclei (AGNs). A cursory look at these spectra reveal XAFS near many photoelectric edges, and hence dust and/or molecules. Therefore, there is already a wealth of data for studying astrophysical dust properties in the ISM and/or the hotter "near binary" environments using the X-ray technique described in this paper. In combination with X-ray scattering studies which will provide important spatial information on distribution (see, e.g., Xiang et al. 2007, which makes a first successful attempt at considering nonuniform dust distributions along the line of sight toward the X-ray dipper 4U 1624 − 490, and references therein), existing grating observations provide us with great potential to map out dust distribution along different lines of sight, while determining the species-specific quantity and composition along these same path lengths to build upon studies in other wavebands. Given spectra with enough S/N (as true of many of the existing data), the spectral resolution afforded by the Chandra gratings will allow us to (1) definitively separate dust from gas-phase absorption, and (2) give us good expectations for being able to directly determine the chemical composition of dust grains containing oxygen, magnesium, silicon, sulfur, and iron. As such, we expect to be able to make significant progress over the next few years in our understanding of the submicron-size molecules which make up a large fraction of the ISM grains, and/or near-binary/near-AGN environments. For the latter, this would enable a better understanding of the evolutionary histories of these black hole and neutron star systems.

For many astrophysical applications, a better understanding of dust properties and its distribution in interstellar space carries many advantages. As an example, consider the supermassive black hole Sagittarius A* in our own Galactic center. Current estimates of the absorption column to Sagittarius A* between X-ray and NIR studies differ by a factor of ∼2, which may be attributed to unknowns relating to the metallicity, gas-to-dust ratio, or the grain size distribution along the line of sight. These uncertainties affect our ability to accurately use the Fe Kα emission line as a diagnostic of the Sagittarius A* hot accretion flow (see, e.g., Xu et al. 2006) and determine the X-ray spectral slope for observed flares. Metallicity uncertainties can be greatly improved as we understand better the nature of the absorption toward the Galactic center. This can be accommodated by the type of study described in this paper, in large part because the spatial distribution of sources with spectra of sufficient S/N and spectral resolution necessary for the proposed X-ray study of dust encompasses much of the region along and around the Galactic plane and toward the Galactic center. Ideally, we would additionally conduct case studies of the fainter binaries in the central parsecs, or at least as many as possible, as close to the Galactic center as possible (see, e.g., Muno et al. 2008 for latest X-ray binary catalog based on Chandra ACIS-I imaging data). Given the relative X-ray faintness of many of these objects, however, such a comprehensive spectral study in the nearest regions of the Galactic center will have to await missions with higher throughput, and at least equal spectral resolution.

As stated, studies of many astrophysical systems can be greatly enhanced by a better understanding of the line-of-sight absorbing material. With the Chandra gratings, we have entered upon an era where condensed matter theory can be applied to astrophysics studies of dust. Invariably, interstellar environments present us with additional complications not of consideration in a controlled laboratory experiment, so we conclude with a thought problem of relevance to dust studies of interstellar environments.

4.1. A Gedanken Problem: Can We Differentiate Dust Content in Different Locations?

The path length along the line of sight to our illuminating sources, be they X-ray binaries in our own Galaxy or black hole systems in other galaxies, is complex. Consider, for example, the simplified scenario of only two components distributed in some manner along the path. This problem then has two parts: the identification of the chemical species of those two components and the determination of their distribution along the path length. Assuming we can identify the two species, can we distinguish a heterogeneous arrangement (dust of composition a in location A plus dust of composition b found in location B) from a homogeneous arrangement (dust of mixed composition AB distributed along the path)?

The interpretation of the X-ray spectrum passing along this line of sight bears a strong similarity to the well-established technique of XAS as performed at synchrotron facilities. The use of XAS to identify chemical species in a chemically heterogeneous sample is common practice in terrestrial laboratories (Lengke et al. 2006 is but one example among thousands in the XAS literature). The common strategy is to interpret the spectrum from the mixture of species as a linear combination of the spectra from its component species, i.e., ∑ni = 0ciαi, where αi represent the different species and ci are their respective fractional contributions as per Lengke et al. (2006). The best determined relative percentages for the linear combination are derived by fits based on minimal χ2, which we recast for our purposes to be

Equation (8)

where βj and αi,j are the normalized cross sections in the jth bin of respectively the mixture and pure forms of the individual compounds, Δβj and Δαi,j are their respective uncertainties, and n is the associated number of compounds and jth bins to sum over; ci has the condition that ∑ni = 1ci = 1. With careful sample preparation, the fractional content of the component species in the spectrum is directly indicative of their fractional content in the physical sample.

To demonstrate this using iron-bearing species of relevance to the ISM, we performed the following experiment in the controlled laboratory of the synchrotron facility. We prepared two samples containing the common terrestrial iron compounds hematite (α-Fe2O3), ferrosilicate (Fe2SiO4), and lepidocrocite (γ-FeOOH). The first sample contained equal parts by weight of hematite and ferrosilicate and the second contained equal parts by weight of all three materials. Samples for measurement by electron yield at beamline 6.3.1 were prepared as described above and the XAS spectra were measured on both mixtures. The hematite and lepidocrocite were very fine-grained, commercially produced powders. The ferrosilicate was commercially produced, but was ground by agate mortar and pestle from a chunk several millimeters across. Consequently, the ferrosilicate component of each sample was much larger grained than the other two components.

Because equal parts by weight were mixed in our two samples, we might expect the measured spectra to be interpreted by equally weighted fractions of the spectra from the three pure materials. In fact, the nature of the interaction between the X-rays and the morphology of the sample must be considered. Because the ferrosilicate component of each sample was much larger grained than the other components, its surface-area-to-volume ratio was considerably smaller. Because the electron yield measurement is surface sensitive, a much smaller fraction of the iron atoms in the ferrosilicate contributed to the actual measurement than of the other two components.

The data on the two mixtures along with the results of the linear combination analysis are shown in Figure 8. The binary mixture proved to be 16 ± 2% ferrosilicate and 84 ± 2% hematite. As expected from particle size, the ferrosilicate is underrepresented in the measured spectrum. The trinary mixture was 8 ± 2% ferrosilicate, 51 ± 7% hematite, and 41 ± 7% lepidocrocite. The two components of similar particle size are evenly represented, while the ferrosilicate is underrepresented.

Figure 8.

Figure 8. Comparison of the linear combination (red) of the pure compounds against the mixture (black) for the problem considered in Section 4.1. In blue and green are the cross sections of the individual compounds reduced according to their fractional contributions in the linear combination; small energy shifts are allowed. The best determined relative percentages for the linear combination are derived by fits based on minimal χ2 according to Equation (8). This color coding applies to both the linear combination of the binary (top) and trinary (bottom) mixtures.

Standard image High-resolution image

This careful consideration of sample morphology in the synchrotron measurement is directly relevant to the interpretation of the astrophysical X-ray spectra. The astrophysical measurement is akin to a transmission measurement at the synchrotron, while the examples shown in Figure 8 were measured in electron yield. Still, the correct interpretation of the XAS spectra requires consideration of the nature of the interaction of the X-rays with the material through which they pass. A ray interacting with a large particle in the ISM will be lost to the satellite measurement by virtue of being completely absorbed by the particle. Consequently, the satellite measurement will be dominated by the small particles in the ISM as those are the only particles that allow passage of photons for eventual measurement at the satellite.

Finally, we must address the topic of distinguishing the distribution of the chemical species along the path length. By itself, a transmission XAS measurement is incapable of distinguishing between a layered and a homogenized sample. In transmission, an XAS measurement on two powders will be the same whether you prepare the powders separately and stack them in the beam path or mix the powders thoroughly before making the measurement. Similarly, the X-ray satellite measurement by itself cannot address the distribution of the species it is able to identify, if co-located at similar redshifts. However, since an XAFS contribution to a photoelectric edge is scaled according to the optical depth of that particular species, we can reasonably conclude that the compound that best fits the astrospectra is the compound where we are seeing the bulk contributions from, even if there might be lesser (e.g., 10%) contributions from other compounds along the line of sight. Therefore, in combination with other astrophysical studies using photon wavelengths that interact differently with the ISM (e.g., IR), the X-ray measurement can play an important role in the full determination of the composition of the ISM.

5. XAFS SCIENCE: PRESENT AND FUTURE

The X-ray energy band has been slow to be exploited for dust studies largely due to instrumental requirements for both good spectral resolution (R ≳ 1000), and throughput. Yet, condensed matter astrophysics (i.e., the merging of high-energy condensed matter and astrophysics techniques for X-ray studies of dust) as a new subfield can be realized in the present era of Chandra and XMM-Newton. As such, X-rays should be considered a powerful and viable new resource for delving into a relatively unexplored regime for determining dust properties: composition, quantity, and distribution. Present-day studies with extant satellites will set the foundation for future studies with larger, more powerful missions such as the joint ESA–JAXA–NASA International X-ray Observatory9 (IXO). Proposed spectral instruments such as the Critical-Angle Transmission grating10,11 (CAT; Flanagan et al. 2007) and Off-Plane Reflection Gratings (Lillie et al. 2007) are designed to provide IXO's baseline resolving power of 3000 under 1 keV, and configurations are contemplated, which will boost this to 5000 or more. Similar baseline resolving power is also expected from the calorimeters for the hard (>6 keV) spectral region. IXO's target spectral resolution R = 3000 in the 0.2–10 keV bandpass, in combination with planned higher throughput (10 times XMM, and 60 times Chandra), will allow us in the future to move beyond the realm discussed in this paper to being able to use XAFS to recreate the crystalline structure of interstellar grains and determine precise oxidation states. In the interest of providing information for the planning of future missions, we provide spectral resolution goals that are mapped to XAFS science hurdles, for gratings and calorimeters (Table 2).

Table 2. X-ray Instrument Capabilities and Goals for ISM Grain Physics Studies

  Gratings 0.25–6 keV FWHM Calorimeter 6–10 keV FWHM  
Spectral Resolutiona R < 500 Chandra R = 1000 Suzakub R = 923 SXS prototypec R ∼ 1500 IDEALd R
Distinguish gas from dust difficulte yesf yes yes  
Differentiate dust not possible yesg okg yesg 3000h
Discern oxides not possible no no maybei 5000

Notes. aSpectral resolution R = E/ΔE = λ/Δλ referenced to 1 keV for the gratings and at 6 keV for the calorimeters. bThe spectral resolution of the calorimeter on board Suzaku which has since failed in-flight. cReasonable expectation for spectral resolution to be achieved with SXS prototype detectors (HgCdTe absorbers and implanted silicon thermometers) planned for the Astro-H mission. The 4 eV resolution (R = 1500) at 6 keV has been achieved in laboratory measurements (Astro-H Team 2008, private communication). The IXO calorimeter baseline performance is 2.5 eV. dApplies to the 0.25–10 keV range for both grating and calorimeter instrument considerations. eInsufficient spectral resolution limits our ability to discern the component of the photoelectric edge due to dust versus gas. fAs stated in Section 1, the soft X-ray band is also complicated by ionized absorption lines imprinted by the hot plasma environment of the illuminating source (e.g., black hole or neutron star) that needs to be modeled out in order to isolate the XAFS features. gPossible with adequate statistics obtained from either long observation time or large detector area. hDifferent forms of iron (silicates, oxides, metal) distinguished, but not different oxide forms. iChallenging measurement requiring measurement statistics approaching the level of synchrotron experiment

Download table as:  ASCIITypeset image

We acknowledge Eric Gullickson, Pannu Nachimuthu, and Elke Arenholz for beamline support. We thank Claude Canizares and Alex Dalgarno for advice and conversations, and Fred Baganoff for discussions relating to Sgr A*. The Advanced Light Source and JKB is supported by the Director, Office of Science, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract no. DE-AC02-05CH11231. J.C.L. is grateful to Chandra grant SAO AR8-9007 and the Harvard Faculty of Arts and Sciences for financial support.

Footnotes

Please wait… references are loading.
10.1088/0004-637X/702/2/970