Brought to you by:

Spectral Deconvolution of the 6196 and 6614 Å Diffuse Interstellar Bands Supports a Common-carrier Origin

, , , and

Published 2018 June 5 © 2018. The American Astronomical Society. All rights reserved.
, , Citation L. S. Bernstein et al 2018 ApJ 859 174 DOI 10.3847/1538-4357/aabd85

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/859/2/174

Abstract

We explore the common-carrier hypothesis for the 6196 and 6614 Å diffuse interstellar bands (DIBs). The observed DIB spectra are sharpened using a spectral deconvolution algorithm. This reveals finer spectral features that provide tighter constraints on candidate carriers. We analyze a deconvolved λ6614 DIB spectrum and derive spectroscopic constants that are then used to model the λ6196 spectra. The common-carrier spectroscopic constants enable quantitative fits to the contrasting λ6196 and λ6614 spectra from two sightlines. Highlights of our analysis include (1) sharp cutoffs for the maximum values of the rotational quantum numbers, Jmax = Kmax, (2) the λ6614 DIB consisting of a doublet and a red-tail component arising from different carriers, (3) the λ6614 doublet and λ6196 DIBs sharing a common carrier, (4) the contrasting shapes of the λ6614 doublet and λ6196 DIBs arising from different vibration–rotation Coriolis coupling constants that originate from transitions from a common ground state to different upper electronic state degenerate vibrational levels, and (5) the different widths of the two DIBs arising from different effective rotational temperatures associated with principal rotational axes that are parallel and perpendicular to the highest-order symmetry axis. The analysis results suggest a puckered oblate symmetric top carrier with a dipole moment aligned with the highest-order symmetry axis. An example candidate carrier consistent with these specifications is corannulene (C20H10), or one of its symmetric ionic or dehydrogenated forms, whose rotational constants are comparable to those obtained from spectral modeling of the DIB profiles.

Export citation and abstract BibTeX RIS

1. Introduction

Nearly 100 years have elapsed since the discovery of the diffuse interstellar bands (DIBs; Heger 1922; McCall & Griffin 2013), comprising over 500 weak, predominantly visible, absorption features seen against distant bright stars (see reviews by Herbig 1995; Tielens & Snow 1995; Fulara & Krelowski 2000; Sarre 2006; Cami & Cox 2014; Geballe 2016). We are still struggling to unambiguously identify even a few of the DIB carriers (Oka & McCall 2011). Recently, ${{{\rm{C}}}_{60}}^{+}$ emerged as a strong candidate for several near-infrared DIBs (Campbell et al. 2015; Campbell & Maier 2017); however, this identification is still being discussed (Galazutdinov et al. 2017; Walker et al. 2017; Galazutdinov & Krelowski 2017). It is widely accepted that the carriers are gas-phase molecules, but there is no consensus on their chemical composition. The more frequently invoked carriers include the neutral, ionic, and dehydrogenated forms of small and large polycyclic aromatic hydrocarbons (PAHs), hydrocarbon chains, and fullerenes. Much of what we know about the carriers has been inferred from survey studies in which a large number of DIBs have been observed against stars at different distances from the Earth (Jenniskens & Desert 1994; Galazutdinov et al. 2000; Hobbs et al. 2008, 2009; Munari et al. 2008; Cox et al. 2014; Baron et al. 2015; Puspitarini et al. 2015; Elyajouri et al. 2017; Law et al. 2017).

The surveys have facilitated correlation studies in which the spectrally integrated absorption strength of a DIB, its equivalent width (EW), can be correlated with that for other DIBs as well as known atomic and molecular absorption features. One of the more interesting outcomes from the correlation studies is the hypothesis of a unique carrier for virtually every DIB (Krełowski et al. 2016). This follows from analyses indicating that only a handful of the more than 5002 = 2.5 × 105 DIB pairs exhibit a very strong correlation, as would be expected from two DIBs arising from the same carrier. Moreover, there is only a single DIB pair, λ6196 and λ6614, that exhibits a "nearly perfect correlation" with a Pearson correlation coefficient of around rP = 0.97 (Moutou et al. 1999; Galazutdinov et al. 2002; McCall et al. 2010; Krełowski et al. 2016). We consider the contrasting hypothesis that there may be many strongly correlated pairs whose correlations are obscured by signal-to-noise considerations, DIB blending, and isotope effects.

The primary focus of this paper is on using spectral profile modeling to explore whether or not the λ6196 and λ6614 DIBs could originate from a common molecular carrier. There are numerous spectral modeling studies of λ6614 (see Bernstein et al. 2015b and references therein; Marshall et al. 2015), several on λ6196 (Walker et al. 2001; Webster 2004), and only one recent attempt to seek a common carrier (Glinski & Eller 2016). The challenge of finding a common carrier is highlighted in Figure 1, where it is evident that the shapes and widths of the two DIBs are very different. The profile differences suggest two different, but closely related, carriers whose correlation could arise from membership in the same DIB family (McCall et al. 2010; Krełowski et al. 2016). As demonstrated by Glinski & Eller (2016), it is challenging to find plausible spectroscopic constants for a single molecule that can quantitatively account for the contrasting shapes and widths of the two DIBs.

Figure 1.

Figure 1. Comparison of the contrasting spectral profiles for the λ6614 (red curve) and λ6196 (blue curve) DIBs for the sightline toward HD 145502 (Galazutdinov et al. 2002). The spectra are displayed in wavenumbers, cm−1, relative to their band origins to facilitate a quantitative shape comparison. We note the very different spectral profiles for these DIBs. The width of λ6196 is approximately half that of λ6614. λ6614 exhibits distinct P, Q, and R branches and an extended red tail. Later, we associate the P, Q, and R features and the red tail with different molecular carriers. The λ6614 RTail feature at −1.5 cm−1 is assigned as the R branch bandhead of the red-tail carrier. λ6196 exhibits distinct P and R branches and does not display a Q branch or a red tail. However, λ6196 may have a Q branch that is obscured by blending with the nearby P and R branches.

Standard image High-resolution image

As a helpful point of reference for this paper, Figure 2 previews some key properties of our prototypical common carrier for the λ6614 and λ6196 DIBs. Corannulene (C20H10) is depicted because it is representative of the deduced shape, symmetry, and size specifications of the carrier. Later, we discuss that a small nearly spherical fullerane, such as C20H, also satisfies the carrier requirements. The carrier is a puckered oblate symmetric top comprised of two equal moments of inertia, IA = IB, perpendicular to, and one moment of inertia, IC, parallel to the highest-order symmetry axis. Because the carrier is puckered, it can have a permanent dipole moment aligned with the highest-order symmetry axis. The distribution of rotational levels for rotation about the different axes can be approximated in terms of two different effective rotational temperatures Trot,C, and Trot,A = Trot,B. The symmetry of the electronic transition moment, parallel or perpendicular, giving rise to a DIB absorption band controls the shape and width of the band. We find that a perpendicular transition moment is required for both λ6614 and λ6196. The above specifications for the common carrier were not assumed a priori. They are the end result of the spectral modeling and analysis of the observed λ6614 and λ6196 spectra presented in this paper.

Figure 2.

Figure 2. Overview of some key properties associated with a prototypical compact PAH common carrier for the λ6614 and λ6196 DIBs. Shown in the upper-right corner is an alternative carrier type, a nearly spherical fullerane. Both carrier types are oblate symmetric tops with a dipole moment along the highest-order symmetry axis.

Standard image High-resolution image

We describe a novel spectral analysis and modeling approach that starts with applying a spectral deconvolution algorithm (Bernstein et al. 2017) to one of the observed λ6614 DIB spectra. This sharpens the observed spectrum and reveals finer spectral details of the carrier(s) that are not evident in the observed data. In general, there are multiple sources of broadening, including lifetime (i.e., Heisenberg uncertainty), vibrational anharmonicity, Doppler, and sensor resolution. We find that the last two are the most important for the DIBs under consideration. We have recently applied the same spectral deconvolution and spectral modeling approach to the 11.2 μm unidentified infrared band (UIR) (Bernstein et al. 2017) where we found that a small fullerene, C24, was a plausible carrier. The details revealed in a deconvolved DIB profile provide tighter constraints on the spectroscopic parameters used to model the data.

Our spectral profile fitting approach includes three parameters not generally considered in the modeling of DIB profiles. They are briefly introduced and discussed in more detail later. The first are sharp cutoffs for the maximum values of the rotational quantum numbers, Jmax = Kmax. This is required to fit the steep declines of the outer edges of the P and R branches of the λ6614 profile. The need for sharp quantum number cutoffs, first introduced in Bernstein et al. (2015b), is more apparent in this work because spectral deconvolution increases the slope of an edge. The second is different effective rotational temperatures for rotations perpendicular and parallel to the highest molecular symmetry axis. The two-rotational temperature model follows from our identification of the carrier as a symmetric top molecule with a dipole moment aligned with the highest-order symmetry axis. Glinski & Eller (2016) suggested that two different rotational temperatures were a necessary component of a viable common-carrier spectral model. The third is the coupling between the vibrational angular momentum from a degenerate vibrational mode and the rotational angular momentum, which is quantified through the Coriolis constant, −1 ≤ ζ ≤ 1 (Herzberg 1989, 1966). Kerr et al. (1996) first introduced this parameter into the spectral analysis of λ6614. The shape and width of a spectral profile are sensitive to the value of ζ. Since the value of ζ is mode specific, transitions terminating on different upper electronic state degenerate vibrational modes can exhibit different shapes and widths, which can account for the contrasting profiles of the λ6614 and λ6196 DIBs.

This paper is organized as follows. We first introduce the observational data used in the study. Then, we apply spectral deconvolution to the λ6614 DIB for HD 145502. Next, we use a carrier-impartial analytical model to fit the deconvolved spectrum in order to untangle the multicomponent spectrum. This provides an estimate of the dominant spectral component to which we apply a physics-based spectral profile model to retrieve estimated spectroscopic parameters for the molecular carrier. These parameters are perturbed in order to fit the observational λ6614 DIB profiles toward HD 145502 and HD 179406. The λ6614 spectroscopic parameters are subsequently applied to the spectral fitting of the λ6196 profiles for the two sightlines. A broad range of related topics are covered in the discussion section, including the (1) overview of the origins and effects of rotational non-equilibrium, (2) physical origin of the doublet for λ6614, (3) observational evidence for the two-temperature model, (4) origin of the extended red tail for λ6614, (5) improving the λ6196–λ6614 correlation, (6) the number of DIB carriers, and (7) candidate carriers. Recommendations are presented for further modeling and observational studies to evaluate the plausibility of the common-carrier hypothesis. We conclude with a summary of key findings. Four appendices which provide detailed spectral modeling and analyses that support the key ideas presented in the discussion section are included.

2. λ6614 and λ6196 Observational Data

The observational data used in this study are presented in Figure 3. The observations were performed at the La Silla observatory in Chile using the Coude Echelle Spectrograph (CES) mounted on the ESO 3.6 m telescope (Galazutdinov et al. 2002). The measurements were obtained at a spectral resolution of R ∼ 220,000, corresponding to ∼0.07 cm−1 for the λ6614 and λ6196 DIBs. The measurement signal-to-noise ratios (S/Ns) were high, ranging over S/N ∼ 600–1000. The S/Ns for the extracted DIB absorption spectra are about an order of magnitude smaller because their maximum absorption depths are on the order of 5%–10% of the full data span. An important criterion for the selected sightlines was that they displayed narrow, predominantly single-peaked velocity distributions, as determined from their Na and K atomic line absorption profiles. This optimized the potential to observe narrow spectral structures inherent to the DIB carriers, which do not arise from a multipeaked velocity profile.

Figure 3.

Figure 3. Comparison of the observed spectra for the λ6614 and λ6196 DIBs for two sightlines, HD 145502 and HD 179402 (Galazutdinov et al. 2002). The vertical red lines facilitate comparison of the locations of the peak features between sightlines. The band origins were taken as 6613.54 and 6196.99 Å for λ6614 and λ6196, respectively.

Standard image High-resolution image

Measurements were performed for seven sightlines. We consider the two sightlines toward HD 145502 and HD 179406 because they are representative of the range of spectral profiles typically observed for these DIBs. There is a notable exception, discussed later, toward Herschel 36 (Dahlstrom et al. 2013; Oka et al. 2013), which displays a highly extended red tail for λ6614 but not for λ6196.

Apart from the obvious differences, noted earlier, in the shapes and widths of the two DIBs for a single sightline, there are also noteworthy differences between the two sightlines. The sightline variability of λ6614 is much less than that for λ6196. For λ6614, there is a slight narrowing of the P–R branch separation and a slight increase in the ratio of the P and R branch peaks to the Q branch peak between HD 145502 and HD 179406. These changes are indicative of a decrease in the effective Trot for λ6614 from HD 145502 to HD 179406. The relative contribution of the red-tail feature is larger in HD 145502, and it displays a distinct peak, denoted as RTail. For λ6196, there are significant differences in the widths and shapes between the two sightlines. These differences are also consistent with the change in the effective Trot for λ6196.

3. Spectral Deconvolution and Modeling of the λ6614 DIB

3.1. Spectral Deconvolution of λ6614 for HD 145502

The spectral deconvolutions and fits for all of the data used Gaussian broadening with widths derived from fits to the observed Na or K atomic line velocity profiles for the DIB sightlines (Galazutdinov et al. 2002). HD 145502 displays a bimodal Na velocity distribution. We based the width estimate on the smaller profile in order to minimize the impact of line saturation effects (Galazutdinov et al. 2002). The EW of the smaller profile is approximately 16% that of the primary profile; however, saturation of the primary feature results in an overestimate of the relative Na abundance for the smaller feature. This is borne out later, where spectral fits show that the smaller feature corresponds to a relative contribution of 6% to the HD 145502 DIB spectra. For HD 179406, we also considered the measured CH+ velocity profile (Kazmierczak et al. 2009), whose width is approximately 15% larger than that for the K line profile. The average of the two measured widths was used in this analysis. The values of the widths are presented in Table 1. The widths include a small, ∼0.01 cm−1, contribution for the sensor resolution in which the Doppler velocity and sensor widths are combined in quadrature. Additional broadening sources could be included in the analysis, such as lifetime and vibrational anharmonicity broadening; however, as discussed below, there is no indication of a significant broadening residual in the deconvolved data.

Table 1.  Molecular Spectral Fit Parameters to the Observational Data Determined Using PGOPHER

Observation B(cm−1)a ΔB/B(%) C(cm−1) ΔC/C(%) ${T}_{\mathrm{rot}}$(K) ${J}_{\max }={K}_{\max }$ ζ ${\gamma }_{G}({\mathrm{cm}}^{-1})$ b
HD 145502 λ6614c 0.0232 −0.70 0.0096 0.0 50 26 −0.60 0.30
HD 179406 λ6614c 0.0232 −0.70 0.0096 0.0 18 26 −0.60 0.28
HD 145502 λ6196 0.0232 1.7 0.0096 0.0 7.0 26 0.62 0.32
HD 179406 λ6196 0.0232 1.7 0.0096 0.0 3.5 26 0.62 0.30
HD 145502 λ6614d 0.0232 −0.70 0.0213 0.0 150 27 0.30 0.30
HD 145502 λ6196d 0.0232 2.6 0.0213 0.0 7.0 27 0.82 0.32

Notes.

aB is the rotational constant for the lower transition level. bFWHM. cThe splitting, Δν, and weights of the doublet components for λ6614 are defined in Figures 7 and 8. dNearly spherical, slightly oblate symmetric top fullerane carrier.

Download table as:  ASCIITypeset image

The deconvolved spectrum for λ6614 from HD 145502 is presented in Figure 4. We see two slightly separated sets of similar PQR branch features, which is suggestive of doublet structure. There also appear to be comparably intense P and R branches with similarly steep outer edges. The overall shape separates into two, slightly overlapping, main regions, an approximately symmetric central PQR region spanning approximately −1.3 to 1.3 cm−1, and an asymmetric red tail below approximately −1.3 cm−1. In contrast to the symmetric region, the red tail does not show evidence of a doublet structure. We interpret this to mean that two blended DIBs, one for the doublet and the other for the red tail, contribute to λ6614. This interpretation is supported below through a detailed spectral analysis.

Figure 4.

Figure 4. Spectral deconvolution of the HD 145502 λ6614 spectrum using a Gaussian of width γG = 0.30 cm−1 (FWHM). The two sets of PQR labeled features suggest a doublet with a separation of about Δν = 0.35 cm−1. In contrast, a doublet structure is not evident in the red-tail feature below −1.3 cm−1. The red-tail feature displays an RTail branch bandhead and an extended red tail. The thin blue and red lines denote the approximate separation of the PQR and red-tail features, respectively, into distinct spectra associated with different carriers.

Standard image High-resolution image

3.2. Analytic Model for the Deconvolved λ6614 Spectrum

Rather than trying to simultaneously model the three overlapping components of λ6614, the doublet and red-tail features, we found it useful to first mathematically disentangle the components using simple analytic functions. The approximate spectroscopic constants derived from subsequent spectral fitting of the analytically separated components provided a good parameter initialization for use in spectral fitting of the observed spectrum. For the doublet feature, we assumed that it was composed of two slightly offset spectra with different weights but identical shapes. The assumption of identical shapes is a convenient simplification whose validity is explored in Section 5.2. For the analytic fit, we represented the P, Q, and R branch features of a doublet component as a combination of a Gaussian and a Lorentzian function. The resulting analytic fit to the deconvolved spectrum is shown in Figure 5. The quality of the analytic fit supports the supposition that a doublet feature is a major constituent of the λ6614 spectrum.

Figure 5.

Figure 5. Application of an analytic spectral fitting model to the deconvolved λ6614 spectrum in Figure 4. The fit parameter values for each branch of the larger doublet component (thin blue line) follow and are denoted by (ν, γG, γL, α), where ν is peak frequency (cm−1), γG and γL are the Gaussian and Lorentzian widths (cm−1 at FWHM), and α is relative peak height. The values are P branch (−0.76, 0.32, 0.32, 0.82), Q branch (0.015, 0.30, 0.30, 1.10), and R branch (0.79, 0.32, 0.32, 0.82). The Gaussian and Lorentzian fits for each branch were normalized to the same area (i.e., equal equivalent widths). We assume both doublet components have the same shape. Thus, except for the peak frequencies, the other parameter values for the smaller component (dashed blue line) are identical to those for the larger component. The doublet splitting is Δν = 0.35 cm−1, and relative weights for the doublet components are w1 = 0.59 and w2 = 0.41. The red-tail residual was determined by subtracting the total analytic fit (thick blue line) from the deconvolved data (black line).

Standard image High-resolution image

Figure 5 also shows that the difference between the deconvolved spectrum and the analytic model defines a residual red-tail feature with an R branch bandhead and a gradually declining red tail. This is an often-encountered DIB shape and is evidence of a separate carrier. While the primary focus of this work is on λ6614 and λ6196, we discuss the implications of the red tail, how it influences the λ6614–λ6196 correlation, and its interpretation as a separate DIB in Sections 5.4 and 5.5.

The analytic model fit to the deconvolved data confirms our previous analysis of the broadened λ6614 spectrum (Bernstein et al. 2015b), which was performed without the benefit of spectral deconvolution. In the prior analysis, we determined a doublet splitting of Δν = 0.25 cm−1 and relative weights for the doublet components of w1 = 0.59 and w2 = 0.41 (i.e., w1 + w2 = 1).

3.3. Spectral Model for λ6614

We used a physics-based spectral model, PGOPHER (Western 2017), to fit the larger analytic component for λ6614. This provided good estimates of the spectroscopic parameters for subsequent use in modeling the observed spectrum. The spectroscopic parameters consist of the lower electronic state rotational constants C (for rotations about the molecular symmetry axis, which is aligned with the dipole axis in Figure 2) and B (for rotations about an axis perpendicular to the symmetry axis); the difference between the upper and lower state rotational constants, ΔB and ΔC; two effective rotational temperatures, Trot; for the tumbling and spinning top motions, the maximum rotational quantum numbers, Jmax and Kmax; the Coriolis coupling constant, ζ; the Gaussian line width, γG; the relative weights of the two components, w2 and w1 (where w1 > w2 and w1 + w2 = 1); and the doublet splitting, Δν. The rotational constants, B and C, are the key to identifying potential carriers because they tightly constrain the shape and size of the carrier (Bernstein et al. 2017).

With the exception of Jmax and Kmax, the above set of spectroscopic parameters represents the minimum number of free parameters required to model the deconvolved doublet spectrum. Assuming a symmetric top carrier, the minimal set of parameters required to model one of the doublet spectral profiles includes B, C, ΔB, ΔC, two Trot, and at least one broadening parameter, γG, to account for Doppler broadening. If sightline atomic line absorption velocity profiles have been measured, as is the case here, then γG is well determined from the profile width and, thus, is not a free parameter. Lifetime broadening, which is described by the Lorentz function, may also be required if the upper electronic state of the transition is very short lived. A doublet spectrum requires two additional parameters, Δν for the separation of the doublet components and the relative weight of one of the components (i.e., just one parameter needed because of the constraint w1 + w2 = 1). This set of parameters applies to a transition with A → A symmetry. If the transition has A → E symmetry, then an additional parameter is needed, the Coriolis coupling constant, ζ. We will see later that the Coriolis parameter is crucial to demonstrating a common-carrier origin for λ6614 and λ6196. A sharp cutoff for the rotational quantum numbers is not typically encountered. However, the narrow and sharply truncated P and R branches for λ6614 (see one of the analytic profiles in Figure 5) indicate the need for this parameter.

We screened a variety of candidate carrier classes comprising different molecular geometries (i.e., linear, spherical, oblate, and prolate symmetric tops) and transition symmetries (i.e., parallel and perpendicular, degenerate and nondegenerate vibrational states). For each geometry and transition symmetry selection, we performed a coarse spectroscopic parameter search, using the above parameter set, to ascertain if a good common-carrier spectral fit to both DIBs was feasible. We found only one combination with the potential to be quantitatively consistent with the common-carrier hypothesis. The screening suggested that the carrier is a dipolar oblate top, with a perpendicular electronic transition symmetry from a nondegenerate into a doubly degenerate vibrational state (this applies to both λ6614 and λ6196). Two classes of carbonaceous molecules are consistent with these specifications, a puckered disk-like compact PAH (e.g., corannulene, C20H10) and a nearly spherical fullerane (e.g., C20H). Not all fulleranes satisfy the symmetric top requirement. The candidate fulleranes that satisfy the derived symmetry and size requirements are C20H and C28H.

Although a compact PAH and a fullerane have very different shapes, from the perspective of spectral profile fitting, they are indistinguishable. The equivalence arises from the Coriolis effect, which manifests in the transition frequencies as (1 – ζ)C (see Appendix B for the detailed spectroscopic origin of this term). Since the transition frequencies are determined by the product of two factors, any combination of factors that yield the same product will exhibit the same transition frequencies. Later, we discuss that a compact PAH and a fullerane with the same B rotational constant but different C rotational constants can produce essentially equivalent spectral profiles for λ6614 and λ6196. Hence, because of the spectroscopic equivalence, we frame our analysis and discussion primarily in terms of a compact PAH. All of the basic concepts and general results found for a compact PAH apply to a fullerane.

The PGOPHER spectral model fit to the HD 145502 λ6614 analytic profile is shown in Figure 6. The fit is good, but not exact. This is not surprising, since the analytic model is not a physics-based model. The sensitivity of the spectral fit to variations in the spectroscopic parameters is depicted in Figures 1521 in Appendix A. There are three noteworthy aspects to the spectral fit. The first is the sharp cutoff at relatively small rotational quantum numbers, Jmax = Kmax = 26 (i.e., Erot(Jmax) ≪ kTrot). When thermodynamic equilibrium is maintained through collisions, a P or R branch exhibits a much broader spectrum, by about an order of magnitude, than a Q branch. Without a J, K cutoff, the calculated P and R branches would peak at J, K ∼ 50 and would extend out to J, K ∼ 150 (see Figure 16).

Figure 6.

Figure 6. PGOPHER spectral model fit to one of the identically shaped doublet components from the analytic fit to the deconvolved data (the thin blue line in Figure 5). The spectral fit parameters are B = 0.0232 cm−1, C = 0.096 cm−1, ΔB/B = ΔC/C = 0.0, Trot = 50 K, Jmax = Kmax = 26, and γG = 0.04 cm−1. The small value for the broadening parameter is a consequence of removing the dominant broadening contribution, γG = 0.30 cm−1 from Doppler broadening, via spectral deconvolution.

Standard image High-resolution image

The second aspect is the sensitivity of the simulated shape and width to the Coriolis constant, as demonstrated by Figure 17. Most of the band profile differences between λ6614 and λ6196 can be explained simply by a difference in the values of the Coriolis constant. The remaining difference, primarily in the widths of the bands, is attributed to the difference in effective rotational temperatures for the bands. The third aspect is the small value of the "residual" Gaussian width, γG = 0.04 cm−1, needed for the spectral model fit to the deconvolved spectrum. This means that the width of the observed data can be fully attributed to the combination of velocity broadening, the slightly separated doublet structure, and sensor resolution. As discussed in Section 5.2, this finding supports Webster's (2004) idea that isotopologs of a DIB carrier may produce several, slightly separated DIB bands.

We performed the PGOPHER spectral fit to the HD 145502 observation shown in Figure 7 by perturbing the spectroscopic parameters from the PGOPHER spectral fit to the analytic component. We found that the relative weights of the doublet components change slightly, from w1 = 0.60 and w2 = 0.40 for the analytic model to w1 = 0.68 and w2 = 0.32 for the spectral fit to the data, due to the inclusion of a third, small component arising from the bimodal Na velocity distribution (Galazutdinov et al. 2000). Inclusion of the third component introduced an asymmetry into the relative heights of the P and R branches, which resulted in a change of the vibration–rotation constant from ΔB/B = 0 to ΔB/B = −0.70%.

Figure 7.

Figure 7. PGOPHER spectral fit to the observed λ6614 data for HD 145502. Also shown are the fit components (thin blue lines). The doublet splitting was constrained by the analytic fit, Δν = 0.35 cm−1, and the broadening was constrained by the fit to the Na atomic line, γG = 0.30 cm−1. There is a small fit component, 6% of the total equivalent width of the doublet components, at −0.805 cm−1 that arises from the small, secondary peak in the velocity profile located at −15.0 km s−1 relative to the main peak. The relative weights of the doublet components are w1 = 0.68 and w2 = 0.32. The red-tail residual (red curve) is the difference between the data (black line) and spectral fit (thick blue line). The values of the other fit parameters are given in Table 1.

Standard image High-resolution image

There is a noticeable fit residual corresponding to the very narrow dip in the data at approximately −0.5 cm−1. We do not believe this feature is associated with the λ6614 DIB. It is not evident in the other λ6614 spectra (Galazutdinov et al. 2002), and the width of the feature is equal to that of the sensor spectral resolution, suggesting it is an atomic emission line.

We turn our attention to the spectral fitting of λ6614 for the second sightline, HD 179406. We performed the λ6614 PGOPHER spectral fit for HD 179406, shown in Figure 8, by adjusting the λ6614 spectroscopic fit parameters determined for HD 145502. The relative weights of the doublet components are essentially identical to those for HD 145502, which, as discussed later, indicate nearly identical 13C abundances for the two sightlines.

Figure 8.

Figure 8. PGOPHER spectral model fit to the observed λ6614 data for HD 179406. Also shown are the fit components (thin blue lines). The doublet splitting was constrained by spectral deconvolution (not shown) to Δν = 0.29 cm−1, and the broadening was constrained by the fit to the average of the CH+ and K atomic line widths, γG = 0.28 cm−1. In contrast to HD 145502, the K atomic line profile for HD 179406 is single peaked (Galazutdinov et al. 2002), hence there is no need for a third component. The relative weights of the doublet components are w1 = 0.67 and w2 = 0.33. The red-tail residual (red curve) is the difference between the data (black line) and spectral fit (thick blue line). The values of the other fit parameters are given in Table 1.

Standard image High-resolution image

There are several notable differences in the two observed spectra that are reflected by the differences in their spectroscopic parameters. The HD 179406 spectrum is slightly narrower than that for HD 145502, which is reflected in its lower Trot, 18 versus 50 K. Even though this is a substantial difference in rotational temperatures, its effects on the band shape and width are small because of the low rotational energy cutoff (see Figure 18). The other difference is for the doublet splitting, Δν = 0.29 and 0.35 cm−1 for HD 179406 and HD 145502, respectively. The 0.06 cm−1 difference in the splitting reflects the uncertainty of measuring the separation of two peaks for a sensor spectral resolution of 0.07 cm−1.

4. Spectral Deconvolution and Modeling of the λ6196 DIB

The overall analysis approach for λ6196 paralleled that for λ6614. We first performed a spectral deconvolution on the observed λ6196 HD 145502 spectrum and then used PGOPHER to derive spectroscopic parameters by fitting the deconvolved spectrum (see Figure 9). For consistency with the common-carrier assumption, several of the spectroscopic parameters used to fit the deconvolved λ6196 spectrum were constrained to be identical to those determined from fitting the λ6614 spectrum. The constrained parameters included B, C, Jmax = Kmax, and γG (adjusted for the difference in the λ6196 and λ6614 band origins). The other spectroscopic parameters were allowed to vary, as is appropriate for transitions into different vibronic excited states. The spectroscopic fit parameters derived from the deconvolved spectrum were applied without modification to model the observed λ6196 spectrum for HD 145502 (see Figure 10). The Doppler broadening, absent in the deconvolved spectrum, was included in the fit to the observed data. The λ6196 spectroscopic parameters for HD 145502 were used to model the λ6196 spectrum for HD 179406, allowing for adjustments to the sightline-dependent parameters, Trot and γG.

Figure 9.

Figure 9. PGOPHER spectral fit to the deconvolved spectrum for the HD 145502 λ6196 DIB. The deconvolution was performed with a Gaussian width of γG = 0.32 cm−1, as deduced from the Na atomic line velocity profile. The fit parameters are presented in Table 1. The value of the broadening parameter in Table 1, γG = 0.32 cm−1, refers to the observed data fit; a smaller value, γG = 0.16 cm−1, was needed to match the deconvolved data. The dominant features in this spectrum are the P and R branches. There is no evidence of a significant red or blue tail or doublet structure.

Standard image High-resolution image
Figure 10.

Figure 10. PGOPHER spectral model fit to the observed λ6196 data for HD 145502. This fit was obtained by applying a Gaussian broadening of γG = 0.32 cm−1 (FWHM) to the spectral model fit for the deconvolved spectrum in Figure 9. We also included a small contribution for the secondary peak in the Na velocity profile for the HD 145502 sightline. The relative magnitude of the small feature was constrained to be 6%, as determined from fitting the λ6614 spectrum (see Figure 7).

Standard image High-resolution image

The spectral deconvolution of λ6196 for HD 145502 (see Figure 9) exhibits a single molecular component, a classic P and R branch spectrum, as opposed to three components, a doublet and a red tail, for λ6614 (see Figure 5). There is no evidence in the deconvolution for either doublet structure or an extended red or blue tail. Thus, there was no need to use an analytic fit to untangle multiple components.

The most significant differences in the spectroscopic parameters for the HD 145502 λ6196 and λ6614 DIBS are (1) the much smaller effective rotational temperature, Trot = 7.0 versus 50 K, (2) the change in sign for the Coriolis constant, ζ = +0.60 versus −0.61, and (3) the much larger "residual" broadening parameter, γG = 0.16 versus 0.04 cm−1. The physical interpretations of the first two differences are discussed later. We suggest that the larger residual broadening parameter for λ6196 may be due to an unresolved doublet structure attributed to isotopologs of the common carrier (see Section 5.2).

The spectral fit to the observed λ6196 spectrum for HD 145502, displayed in Figure 10, was obtained simply by applying the measured velocity broadening to the spectral fit for the deconvolved data. The spectral fit to λ6196 for the other sightline, HD 179406, is shown in Figure 11. It was obtained by modifying the sightline-dependent properties, reducing the rotational temperature, Trot = 3.5 versus 7 K, and the broadening parameter, γG = 0.30 versus 0.32 cm−1, determined for HD 145502. The latter difference reflects the slightly narrow velocity distribution for HD 179406 versus HD 145502 (Galazutdinov et al. 2002). The significance of the very low rotational temperature, close to the cosmic microwave background (CMB) temperature of 2.7 K, is discussed later. We note that a proportionally large temperature difference was determined for λ6614, Trot = 18 versus 50 K, for the two sightlines.

Figure 11.

Figure 11. PGOPHER spectral model fit to the observed λ6196 data for HD 179406. This fit required a significantly lower rotational temperature, Trot = 3.5 vs. 7.0 K, and a slightly smaller Gaussian width, γG  = 0.30 vs. 0.32 cm−1, than were determined for HD 145502.

Standard image High-resolution image

We consider the ramifications of basing the spectral modeling on a fullerane instead of a compact PAH carrier. The retrieved spectroscopic parameters for the HD 145502 λ6614 and λ6196 DIBs for a prototypical fullerane carrier are presented in Table 1. The spectral fits are not presented because they are virtually identical to those already shown for the compact PAH carrier (see Figures 7 and 10). We constrained the B rotational constant and the Gaussian broadening widths to be identical for both carrier types. Only the C rotational constant needed to be varied in order to simulate the compact PAH and fullerane geometries. The other parameters were also allowed to vary, as is appropriate for the different carriers. As mentioned earlier, while the ζ and C parameter values are quite different between the two carrier types, the equivalence between their spectral fits is mainly due to the equivalence of their (1 − ζ)C terms. For λ6614, the values of this term are nearly identical, 0.154 and 0.149 cm−1 for the PAH and fullerane, respectively. Similarly, for λ6196, the values of this term are 0.00365 and 0.00368 cm−1 for the two carrier types. The most striking difference between the parameter values is for the Trot for λ6614, 50 and 150 K for the PAH and fullerane, respectively. While this temperature difference is large, its effect on the spectral profile is minor because of the low rotational energy cutoff (see Figure 18). In summary, from a spectral fitting perspective, both a puckered compact PAH and a fullerane are plausible carriers.

5. Discussion

5.1. Overview of the Origins and Effects of Rotational Non-equilibrium

We summarize here the key physical ideas with regard to the origin and effects of rotational non-equilibrium for symmetric top molecules (i.e., the DIB carriers) and then present a more detailed explanation, in Appendix B, based on the fundamental spectroscopy for symmetric tops. There are three key ideas. The first is that there are two effective rotational temperatures, as originally suggested by Glinski & Eller (2016), that characterize the populations of the rotational states for a polar symmetric top in which the dipole moment is aligned with the highest-order symmetry axis. They correspond to rotations perpendicular to and about the highest-order symmetry axis. The perpendicular rotations can emit and absorb radiation because the dipole moment is changing its orientation and thereby produces an oscillating electric field. Because de-excitation collisions are rare in the low-density ISM, the primary de-excitation pathway for a rotating molecule is through emission (see Appendix C for a quantitative comparison of radiative emission and collision rates). An excited molecule will spontaneously emit its perpendicular rotational energy until it reaches equilibrium with the ambient radiation field, which is frequently at or near the CMB at a temperature of 2.73 K but will never be colder (Bernstein et al. 2015a).

In contrast, rotations about the highest-order symmetry axis do not change the orientation of the dipole moment and therefore cannot readily, via allowed transitions, absorb or emit photons. As a result, the rotational temperature for this motion will fall somewhere between two limits. If the collision rate is zero, then its rotational temperature will reflect its nascent (i.e., initially formed) distribution of rotational states. If the collision rate is nonzero, then the rotational temperature will eventually reach thermal equilibrium with the translational temperature. However, this may take a very long time, and other timescales, such as photodissociation, may interrupt the full thermalization of the rotational states.

Although the two-temperature model has not previously been used to model DIB spectra, it has been applied to the closely related problem of modeling interstellar microwave emission spectra. Loren & Mundy (1984) used the two-temperature approximation, which they referred to as the rotational temperature equilibrium (RTE) method, to analyze the rotational emission spectrum of the symmetric top molecule CH3CN in interstellar molecular clouds. In the two-temperature/RTE model, the relative population between K stacks and the population of J levels within a K stack are described by different effective Trot. The J levels for a given K correspond to different end-over-end tumbling rates for a fixed spin rate about the symmetry axis. The Trot characterizing the J-level population within a K stack is very cold because these J levels can efficiently relax via rotational radiative decay. We expect the Trot for the different K stacks to be nearly the same, as is borne out by Figure 12 in Loren & Mundy's (1984) analysis of CH3CN. The validity of the two-temperature model and the Jmax = Kmax approximation for the prototypical common carrier for the λ6614 and λ6196 DIBs are discussed in Appendix C.

The second key idea is that the value of the Coriolis constant determines which one of the two effective rotational temperatures characterizes a spectrum. For instance, in Appendix B, we show that in the limit of ζ = 1, the spectrum depends only on the very cold Trot associated with the tumbling motion. As a result of the low rotational temperature, the band is narrow. Conversely, for ζ = −1, the spectrum is sensitive to the hotter rotational temperature associated with the spinning motion, and the band is much wider. Examples of synthetic DIB spectra corresponding to the two limits for ζ are presented in Appendix B (see Figure 22). While these are idealized limiting cases, they convey the idea that the effective rotational temperature for a DIB band will depend on its value for ζ, which can be different for each degenerate mode. Thus, two DIBs arising from the same ground state but terminating on different excited degenerate vibrational modes can exhibit very different shapes and widths.

The third key idea is that there is a steep cutoff in the maximum rotational quantum states, which we approximate as a hard cutoff, Jmax = Kmax. The existence of the sharp cutoff is reinforced by the spectral deconvolution of λ6614; however, the physical origin of the cutoff is a matter of conjecture. λ6196 is insensitive to the existence of the sharp cutoff because its low effective Trot renders it insensitive to the higher J, K states associated with the cutoff (see Figure 22). Sharp rotational level cutoffs have been observed in absorption and emission spectra for many diatomic hydrides, with OH as a particularly noteworthy example (Carrington 1964). The cutoff arises from a predissociation barrier in an excited electronic state. J states at or above the predissociation barrier rapidly dissociate, leading to very broad lines with small absorption and emission spectral intensities. Conceptually, predissociation or an analogous transformation (see Appendix D) could occur for a much larger DIB molecular carrier, such as a PAH or fullerane.

In summary, the two-temperature model, the Coriolis effect, and the steep cutoff for the maximum rotational quantum states are necessary and physically credible components of a common-carrier spectral model for the λ6614 and λ6196 DIBs.

5.2. Origin of the λ6614 Doublet

We attribute the doublet to a combination of two isotopic shifts originating from the substitution of 13C for 12C in the carrier. One shift is due to the difference in the zero-point vibrational energies between the upper and lower electronic states (Webster 2004) of a DIB transition. The other is the isotope shift in the fundamental frequency of the assumed singly excited, degenerate vibrational mode in the upper electronic state of the λ6614 DIB transition. All the other vibrations in the upper electronic state are assumed to be in their ground state, as well as all the vibrations in the lower electronic state. We estimate below the shifts and widths of the doublet component arising from one or more 13C substitutions.

Our estimates are based on the prototype carrier corannulene for which NC = 20. A single 13C substitution is considered first, as it is the dominant substitution for typical interstellar 13C abundances. According to Webster (2004), substitution of a single 13C into a carrier with NC = 20 will result in a zero-point energy shift of approximately +0.9 cm−1. This is likely as an upper limit estimate because it rests on the assumption that the promotion of a bonding pi electron into an excited electronic state is equivalent to the complete elimination of a C–C bond. In general, we expect the decrease in electronic stability to be delocalized, resulting in a small reduction in the average bond stiffness in the upper electronic state relative to the lower, ground electronic state. We are not aware of any experimental or accurate first principles determination of the zero-point shift for a large PAH. Regardless of the precise value of the zero-point shift, all DIBs arising from ground vibrational state transitions in the upper and lower electronic states should exhibit blueshifts in their peak locations that will increase with increasing 13C substitution.

If not fully resolved, the 13C spectral component should produce a slight broadening and blue shaded asymmetry to the observed DIB profile. The variable (i.e., depends on the 13C abundance) broadening and asymmetry attributed to the zero-point isotope blueshift may be evident in several blueshifted DIB profiles observed by Galazutdinov et al. (2015).

Unlike the zero-point isotope shift, which involves summing over all of the vibrational modes, the isotope shift in a fundamental frequency depends specifically on which vibrational mode is excited. Since neither the actual carrier nor specific mode is known, we can only establish a plausible range for the shift of the fundamental frequency. This was accomplished using quantum chemistry computations for the fundamental frequencies of corannulene in which a single 13C substitution was considered for the two distinct edge positions and one distinct interior position (see Figure 2). The resulting distribution of shifts for the degenerate modes is shown in Figure 12. Not displayed is the small fraction of shifts, ∼6%, that fall within the range of −5 to −22 cm−1. DIBs arising from such highly shifted modes, if observable, would appear well separated (i.e., not blended) from the λ6196 and λ6614 DIBs. The preponderance of shifts, about 80%, fall within the range of −1.9 to 0.1 cm−1. Combining this range with the zero-point shift implies that the most probable combined shifts fall within the approximate range of −1.0 to +0.8 cm−1. The observed shifts for λ6614 of ∼0.3 cm−1, as well as the lack of an observable shift for λ6196, are consistent with the estimated range of shifts.

Figure 12.

Figure 12. Histogram of the calculated isotope shifts (difference in vibrational frequencies between substituted and nonsubstituted isomers) for the doubly degenerate vibrational modes of corannulene based on the substitution of a single 13C at the three different symmetry positions. Corannulene possesses 84 fundamental vibrational modes of which 68 modes comprise 34 degenerate pairs. The shifts were binned with a width of 0.2 cm−1. Also shown is the cumulative histogram distribution (red curve).

Standard image High-resolution image

Figure 12 shows that ∼2/3 of the redshifts fall within a small range of −0.7 to +0.1 cm−1. If we assume that this fairly probable range encompasses the shifts of the λ6614 and λ6196 DIBs, then this implies that the zero-point shift is considerably smaller than the presumed upper limit estimate of +0.9 cm−1, most likely in the range of ∼0–0.4 cm−1. This reduced range for the zero-point shift makes physical sense. The widths of most DIBs fall within the range of ∼1–2 cm−1 (Hobbs et al. 2008, 2009). If the typical zero-point shift were +0.9 cm−1, then, contrary to observations, many DIBs would exhibit obvious doublet features.

λ6196 does not exhibit distinct doublet components, even after the application of spectral deconvolution (see Figure 9). However, we found that an extra broadening contribution, γG = 0.16 cm−1, was required to fit the deconvolved spectrum. Also based on spectral modeling, Webster (2004) estimated a comparable zero-point broadening contribution of 0.25 cm−1 for λ6196.

We obtained good spectral fits to the λ6614 DIBs based on the simplifying assumption that the doublet components have identical shapes. This assumption results in two additional spectral fitting parameters, the doublet separation, Δν, and the relative weights, ww2 = 1, of the two components. However, one expects that the smaller doublet component, due to the 13C substitutions, should be wider than the larger, pure 12C component. This follows because single and multiple 13C substitutions over the different types of locations within the carrier will have different shifts, thus imparting "extra" width to the effective profile. A third adjustable parameter would be required to account for the extra width of the smaller doublet component. We conjecture that allowing the width to vary would incrementally improve the fits, and consequently, we decided not to include it in our analysis.

The relative weights of the doublet components are consistent with the isotope model. We associate w1 with the pure 12C carrier and w2 with the sum of the weights for all of the 13C substitutions. Table 2 presents the top four fractional isotope abundances for corannulene as a function of the 12C/13C abundance ratio, R. The abundances are based on a Poisson distribution as discussed in Webster (2004; see Equation (8)). For galactic molecular clouds, R ranges from ∼20–70, with an average of ∼45 (Savage et al. 2002). We retrieved values of w1 = 0.67 and 0.68 for HD 145502 and HD 179406, respectively. Table 2 shows that these values for w1 imply R ∼ 50, which is in line with the molecular cloud average of R ∼ 45. The ratio R ∼ 50 means that ∼82% of the 13C carriers correspond to a single 13C substitution, with the remaining ∼18% predominantly due to double 13C substitutions.

Table 2.  Fractional Abundances of Different Numbers of 13C Substitutions, ${f}_{n13C}$, for NC = 20 vs. the Ratio, R, of Interstellar 12C/13C Isotope Abundances

R f0a f1 f2 f3
20 0.368 0.368 0.185 0.0613
45 0.641 0.285 0.0633 0.00991
50 0.670 0.268 0.0536 0.00715
55 0.693 0.253 0.0460 0.00557
60 0.716 0.238 0.0398 0.00442
70 0.751 0.214 0.0307 0.00292

Note.

aw1 = f0 and w2 = f1 + f2 + f3 +....

Download table as:  ASCIITypeset image

At the high end of the molecular cloud 13C abundance range, R ∼ 20, the appearance of the λ6614 spectrum would be dramatically different from those considered here. The majority of the carriers, about 63%, would contain one or more 13C. This would substantially shift the peak locations and broaden the λ6614 spectrum to the red. Additionally, the distinctive PQR branch structure would be difficult to discern.

We would expect that the isotope shifts for the sightlines would be nearly identical since they have essentially the same 13C abundances. However, the shifts differ by 0.06 cm−1. We attribute the difference to the effects of sensor resolution. The accuracy to which the separation of two narrow Q branch peaks can be measured is limited by the measurement spectral resolution of 0.07 cm−1. Variability arises because variations in the sightline radial velocities can result in a variation of up to one detector element in the apparent separation of narrow features.

In summary, isotope shifts are expected to be evident in most DIB spectra, either as separate features or as extra broadening and asymmetry. Redshifts may occur for DIB transitions from a ground vibrational and electronic state to a vibrationally excited upper electronic state. Blueshifts will occur for DIB transitions between ground vibrational levels of the lower and upper electronic states.

5.3. Observational Evidence for the Two-temperature Model

The DIB correlations study of Galazutdinov et al. (2002) supports the two-temperature model. They noted substantial variations of the widths of λ6196 for the sightlines investigated. In contrast, λ6614 exhibited minimal variability in its width. As discussed above, the two-temperature spectral model displays these same characteristics. We showed that the predicted shape and width for λ6614 had minimal sensitivity to Trot due to the sharp cutoff for Jmax (see Figure 18). The predicted spectrum for λ6614 is closely tied (see Appendix B) to the non-emitting rotations about the carrier's highest symmetry axis, implying that the local translational temperature should govern Trot. Thus, even though Trot will vary substantially among different sightlines, this will have little impact on the predicted width of λ6614. In contrast, we attribute the λ6196 spectrum to rotations about the emitting rotational axes whose Trot will be much colder than that for λ6614 and tied closely to the cold Trot measured for smaller polar molecules. The predicted widths for λ6196 are sensitive to small changes in temperature (see Figure 21) because even a variation of 1 K corresponds to a large fractional temperature change for a small Trot.

Kazmierczak et al. (2009) present a more challenging observational result for the two-temperature model. They discovered a significant positive correlation between the width of the λ6196 band and the rotational temperature of a nonpolar C2 molecule. They hypothesized that this correlation implied that the λ6196 carrier is also a nonpolar molecule because the widths for a polar molecule should be nearly invariant for all sightlines due to rotational emission. However, their hypothesis was based on the implied assumption of a single effective rotational temperature for the carrier. Whether or not the two-temperature model for a polar molecule can also exhibit this correlation is an open question. Its resolution requires the development of a first principles theoretical model for the collisional and radiative excitation and de-excitation of a polar symmetric top molecule. This is a challenging endeavor and beyond the scope of this work. However, we anticipate that such a correlation is likely to exist for the two-temperature model. This is because the excitation and de-excitation of rotation about the highest-order symmetry axis are analogous to those for a nonpolar molecule. We expect that the shape and width of any band for a polar symmetric top molecule will depend on both their hot and cold rotational temperatures, and therefore they should exhibit a positive correlation with the rotational temperature for a nonpolar molecule.

5.4. Origin of the Extended Red Tail for λ6614

For a variety of reasons, λ6614 is one of the most studied DIBs. Its prominent PQR-like structure is indicative of a molecular carrier. It has the strongest known correlation with another DIB, λ6196. It is one of just a few DIBs that may also appear in emission in the Red Rectangle (Sarre et al. 1995). It has an extended red tail that extends much farther to the red—i.e., the so-called extended tail to the red (ETR; Dahlstrom et al. 2013; Oka et al. 2013)—than is typically observed in Herschel 36 (see Figures 7 and 8). Understanding the physical origin of the enigmatic red tail is important because it is central to resolving several outstanding issues, such as whether or not λ6614 is a blended DIB and whether or not the λ6614 and λ6196 DIBs arise from the same carrier.

A number of recent studies have explored different origins for the red tail. Several ascribed to the red tail unusually large values of the vibration–rotation constants (Oka et al. 2013; Glinski & Eller 2016). This approach did not produce quantitative fits to the DIB and ETR data for λ6614. Another study attributed the red tail to shifted hot bands of the relatively symmetric central PQR structure (Marshall et al. 2015). The approach was applied to spectra from two sightlines toward HD 147889 and HD 179406, featuring large and small red-tail contributions, respectively. Including hot bands substantially improved the fit for the large tail spectrum. Only a marginal improvement was obtained for the small tail spectrum. This approach has yet to be applied to the ETR and Red Rectangle emission DIBs. A third approach suggested that λ6614 is a blended DIB in which the central symmetric PQR structure and the red tail are due to different carriers (Bernstein et al. 2015b). This approach produced quantitative spectral fits to the DIB, ETR, and Red Rectangle emission spectra.

The enhanced spectral resolution provided by the deconvolution (see Figures 4 and 5) reveals features that are most consistent with the blended DIB hypothesis. The deconvolved spectrum suggests that the central PQR region is comprised of two slightly shifted components. This is confirmed by the analytic spectral fit that shows (1) there are similarly shaped doublet components separated by 0.35 cm−1, and (2) the doublet components have comparable abundances (i.e., w1 = 0.60 and w2 = 0.40). The deconvolution also displays a sharp feature at a frequency of ν ∼ −1.5 cm−1. If this feature was associated with the central components, then one would expect it would also exhibit a second slightly shifted sharp feature. There is no indication of a second component in the deconvolved spectrum. Furthermore, it is apparent that the smoothly declining red-tail residual cannot be quantitatively fit by a linear combination of shifted highly structured central spectral components (i.e., the hot band hypothesis). The red-tail residual has a classic molecular spectral structure, a blue bandhead with an extended red tail, which can be fit with spectroscopic constants very different from those for the PQR feature (Bernstein et al. 2015b). The spectroscopic constants for the red-tail residual also provided quantitative fits to the ETR and Red Rectangle spectra. We conclude that there are different molecular carriers for the red tail and central PQR spectral components; therefore, λ6614 is a blended DIB.

The origin of the red tail also bears strongly on whether or not λ6614 and λ6196 arise from a common carrier. If we assume that the two DIBs share a common carrier, then this argues against the hot band origin of the red tail. The hot bands originate from a vibrationally excited ground electronic state of a DIB transition and therefore they would likely produce comparably strong red tails for all DIBs sharing the same ground state. There is no evidence for a significant red tail in the λ6196 spectra (see Figures 9–11; Galazutdinov et al. 2002, 2008). There is a minor feature that appears slightly to the red of λ6196 in the Herschel 36 data. However, the shape of this feature, a distinct peak and only slight overlap with λ6196 (see Figure 6 in Krelowski 2014), suggests it is a unique DIB, not the red tail of λ6196. Thus, we conclude that the common carrier and blended DIB hypotheses are self-consistent.

5.5. Improving the λ6196–λ6614 Correlation

Krełowski et al. (2016) demonstrated a significant variability of the ratio of EWs for λ6614 and λ6196 that was not due to measurement noise. They interpreted this to mean that there are separate but closely related carriers for the DIBs. We suggest that this "extra" variability arises from uncorrelated spectral features contaminating the DIBs. The uncorrelated features include the red tail for λ6614 and a number of smaller features for λ6196 (see Figure 10). We explore the idea that removing these uncorrelated features should measurably improve the already high correlation between the DIBs.

Our initial exploration of this idea is based on the high-S/N and high-resolution spectral profiles observed by Galazutdinov et al. (2002). The quality and resolution of these data facilitated our effort to separate the blended contributions to these DIBs. We removed the red-tail contribution from the λ6614 spectra by subtracting a normalized template red-tail model determined from our earlier spectral analysis (i.e., the red-tail residual in Figure 8). For each observed spectrum, we normalized the red-tail template to the data at ν = −1.6 cm−1, which is free of contamination from the central PQR feature. For λ6196, we removed the extraneous features using a model fit to each of the deconvolved spectra (see Figure 10). We did not apply this procedure to λ6614 because its much larger EW, about four times that of λ6196, proportionally reduced the contribution of contaminating features. In Figure 13, we plot the EWs of λ6614 against those of λ6196, comparing values based on our modified profiles (using the model spectra for λ6196 and the red-tail subtracted spectra for λ6614) to values based on the observed data. The parameter values used in the plot are presented in Table 3.

Figure 13.

Figure 13. Correlation of the equivalent widths for the λ6614 and λ6196 DIBs. The previously observed and modified (red tail removed) data points (see Table 3) are shown as the black and red squares, respectively. The new and modified points for HD 145502 are shown as the blue and green circles, respectively. Representative error bars are shown for one of the red points. Also shown are the linear fits and the corresponding Pearson correlation coefficient, rP. The revised Pearson correlation coefficients for the new HD 145502 measurements are shown in corresponding colors.

Standard image High-resolution image

Table 3.  Modified Equivalent Widths for the λ6614 and λ6196 DIBs

Source EW6196(mÅ)a ${\mathrm{EW}}_{6196,\mathrm{mod}}$(mÅ)b EW6614(mÅ)a ${\mathrm{EW}}_{\mathrm{Red}\mathrm{Tail}}$ (mÅ)c ${f}_{\mathrm{Red}\mathrm{Tail}}$ d
HD 144757 10.6 10.0 45.3 8.3 0.22
HD 144217 12.1 11.6 50.8 9.7 0.24
HD 144470 12.0 10.3 57.8 13.3 0.30
HD 147165 15.1 12.2 60.2 13.5 0.29
HD 145502 15.8 13.9 57.8 13.3 0.30
HD 145502e 14.6 12.8 59.3 13.6 0.30
HD 184915 16.4 14.8 76.4 13.9 0.22
HD 179406 19.8 19.0 96.8 17.6 0.22

Notes.

aObserved equivalent widths. bModified equivalent width based on λ6196 deconvolution (see text for details). cDerived red-tail equivalent width (see text for details). dFractional contribution of red tail, $f={\mathrm{EW}}_{\mathrm{Red}\mathrm{Tal}}/({\mathrm{EW}}_{6614}\mbox{--}{\mathrm{EW}}_{\mathrm{Red}\mathrm{Tail}}$). eNew observation at Las Campanas.

Download table as:  ASCIITypeset image

The HD 145502 data and modified data points exhibit the largest variance from their respective linear fits. Recently, G. A. Galazutdinov & Y. Beletsky (2017, private communication) reobserved the λ6196 and λ6614 DIBs at the Las Campanas Observatory in Chile using the MIKE spectrograph (R ∼ 80,000) mounted on the Magellan/Clay telescope. The new measurements bring HD 145502 into better agreement with the correlations based on the other data points and significantly improve the Pearson coefficients for both the observed and the modified data, with the former increasing from 0.917 to 0.957 and the latter from 0.935 to 0.970. The new data are at the edges of the one sigma uncertainty bounds of the earlier measurements. Thus, the differences between the new and old measurements are most likely due to measurement statistical fluctuations.

Further support for this conclusion is provided by previous measurements of HD 145502 at the McDonald Observatory in Texas using an echelle spectrograph (R ∼ 60,000) mounted on the Canada–France–Hawaii Telescope (Krelowski & Sneden 1993). Analysis of these data by Wszołek & Godłowski (2003) yielded EWs of 14 ± 1 mÅ for λ6196 and 62 ± 3 mÅ for λ6614. These values agree well with the new measurements discussed above (see Table 3).

From Table 3, we see that there is a significant variability of the red-tail contribution to the EWs for λ6614. The ratio of the red tail and central PQR EWs, fRed Tail, vary by a factor of ∼1.4 from lowest to highest. Other observations, for example, HD 166937 (μ Sgr; Kerr et al. 1996), display a much smaller red-tail contribution, fRed Tail ∼ 0.1, than the observations presented in Table 3. We surmise that the actual variation of the red tail with respect to the central PQR feature is closer to a factor of two, supporting the idea that the red tail and the central PQR features arise from different carriers.

We conclude that the modified data exhibit a significantly improved Pearson correlation over that for the observations (0.970 versus 0.957). However, final confirmation of this initial finding requires testing with a more extensive data set than the seven sightlines considered in this study.

5.6. How Many DIB Carriers?

The number of "perfect" correlations and whether or not they arise from a common carrier has important implications with respect to the total number of DIB carriers. Only a handful (∼11) of strongly correlated DIB pairs, defined by a correlation coefficient of RP ≥ 0.95, have been identified (Xiang et al. 2012). This is surprising since the ∼500 observed DIBs define 1.25 × 105 DIB pairs. The small number of strong correlations has been interpreted to mean that almost every DIB has a unique carrier. This idea seems at odds with the widely held premise that the DIBs arise from polyatomic molecules with a size range of NC ∼ 10–100 carbon atoms. One would expect many of these molecules to exhibit multiple strongly correlated bands. The correlations could arise from Franck–Condon sequences, multiple excited electronic states, and excited-state splitting from a variety of mechanisms, such as a spin–orbit interaction and isotopologs. We argue that there may be many strongly correlated DIBs, but the number of observable strong correlations is severely limited by S/N and DIB blending considerations.

We have formulated a simple statistical model for the number of observable strong correlations, NObs, based on observational properties for the DIBs. The resulting expression is

Equation (1)

where fB is the fraction of blended DIB bands, fEW is the fraction of DIBs whose maximum observed EW has a sufficiently large S/N to enable the determination of a strong correlation, NDIB is the number of known DIBs, and NB is the average number of strongly correlated bands for a molecular carrier. The first three factors are uncorrelated probabilities, and the last term is the total number of DIB pairs. We can estimate reasonable values for all of the quantities of Equation (1). The number of known DIBs is NDIB ∼ 500. The fraction of blended DIBs has previously been estimated to be fB ∼ 0.5 (Bernstein et al. 2015b). Prior studies on DIB correlations have focused on DIB pairs in which each band had a maximum EW in the range of EW ∼ 50–100 mÅ (Xiang et al. 2012). Observational studies of the distribution of DIB EWs suggest that this range of EWs corresponds to fEW ∼ 0.05–0.15 (Hobbs et al. 2008, 2009). We guesstimate that the number of correlated bands per molecule is in the range of NB ∼ 1–10. Predictions for NObs based on these parameter values are presented in Figure 14.

Figure 14.

Figure 14. Estimate of the number of observable strong DIB correlations, NObs, as a function of the number of DIB bands per molecule, NB, and the fraction of DIBs, fEW, with sufficiently high signal-to-noise EWs that can support a high correlation coefficient.

Standard image High-resolution image

We have not included the potential effect of isotopologs on DIB correlation. It is conceivable that an isotopolog may be shifted outside of the narrow spectral range used to compute a DIB's EW. Since the ISM 13C abundance varies significantly, by about a factor of 2, this can add noise to the DIB EW distribution, thereby weakening the correlation.

We use the predictions to estimate the number of unique DIB carriers. If we assume from the observational result that NObs ∼ 11, then Figure 14 implies that for fEW ∼ 0.10–0.15, NB ∼ 5–10. The number of observable correlations is strongly tied to fEW since it enters Equation (1) as ${{f}_{\mathrm{EW}}}^{2}$. For fEW ∼ 0.1, the number of observable correlations is two orders of magnitude fewer than the total number of DIB pairs. This model suggests that the number of unique carriers, NDIB/NB, is approximately ∼50–100. The model also implies that if the limiting effects of DIB blending and S/N could be overcome, then there would be ∼1000 observable strongly correlated DIB pairs.

5.7. Candidate Carriers

The spectral analysis supports two types of oblate, polar, symmetric top carriers: a compact, puckered PAH and a fullerane. Because of the ambiguity between the size-dependent rotational constants and rotational temperatures, the carrier size is not well determined from spectral analysis alone. As discussed in previous work (Bernstein et al. 2015b), the shape of a band depends on the product of the rotational constants and temperature, BTrot, not on their individual values.

We summarize results from analyses presented in Appendices B and C that constrain Trot and thereby enable a determination of B and the associated carrier size. The small end of the size range, which corresponds to the largest B, is limited by the lowest possible rotational temperature, Trot = 2.73 K for the CMB. This temperature is close to the Trot = 3.5 K that we determined for λ6196 for HD 179406 based on B ∼ 0.02 cm−1. This implies a size of NC ∼ 20 carbon atoms. In Bernstein et al. (2015a), we argued that the DIBs and AME (anomalous microwave emission) likely arise from the same collection of carriers. This led to constraints on their temperatures and rotational constants, Trot ∼ 3–10 K and B ∼ 0.01–0.06 cm−1, that are consistent with the findings of this work.

The candidate PAH carrier most consistent with the shape, size, and polarity constraints is corannulene, C20H10. However, we cannot rule out other potential carriers related to circumtridene, C36H12 whose rotational constants are about half of those for corannulene. Both molecules are compact PAHs with a large dipole moment. At present, we consider all of the ionic and symmetric top dehydrogenated variants of these molecules (e.g., symmetrically dehydrogenated C20H5 and fully dehydrogenated C20) as candidate carriers. All of the candidate carriers have multiple degenerate vibrational modes. A key consideration is whether or not they have vibronic transitions that can give rise to λ6614 and λ6196.

With the exception of corannulene (Hardy et al. 2017), quantitative information on the excited electronic states of these species required for this evaluation is not yet available. Rouille et al. (2008) measured the ultraviolet-visible laboratory spectrum of corannulene and did not observe any absorption bands in the visible region. Corannulene has been proposed as a potentially abundant interstellar molecule (Thaddeus 2006) readily observable in the microwave spectral region because of its large dipole moment. Several attempts to detect interstellar corannulene in the microwave region were unsuccessful (Thaddeus 2006; Pilleri et al. 2009). We cannot comment on the results of Thaddeus (2006) because the details of the analysis were not published. However, Pilleri et al. (2009) did not account for the fast radiative decay at the high J = 135 state used for their observation. The radiative decay rate is about four orders of magnitude faster than a typical ISM collision rate (see Appendix C). This disparity in the radiative decay and collision rates would result in a negligible population of and negligible emission from the J = 135 state. Hence, it is premature to rule out the possibility of a significant abundance of corannulene based on the Pilleri et al. (2009) measurements.

Two candidate fullerane carriers, including their neutral and ionic forms, are also consistent with the size, shape, and polarity constraints, C20H and C28H. We expect these molecules to be slightly oblate symmetric tops with average rotational constants of approximately 0.023 and 0.013 cm−1, respectively. While the modeled fullerane and compact PAH carriers have different shapes, they are equivalent from a spectral fitting perspective. They have comparable B values (see Table 1) and nearly equivalent values of (1 − ζ)C (see Appendix B). However, we presume that all variants of the fullerane and PAH candidate carriers have unique excited electronic state energies, and that only one (or no) carrier will have transition energies consistent with the λ6614 and λ6196 DIBS.

5.8. Recommendations for Future Efforts

Recommendations for future investigations to clarify the existence and number of DIBs originating from a common molecular carrier include the following:

  • 1.  
    Apply the λ6614–λ6196 correlation improvements to a more statistically meaningful data set (e.g., Krełowski et al. 2016). As noted in the introduction, the Pearson correlation for a number of larger data sets typically falls around rP ∼ 0.97. Assuming the ∼0.02 correlation improvement found for the small data set considered here, this implies an improved correlation of rP ∼ 0.99. Such a strong correlation would substantially strengthen the case for a common carrier.
  • 2.  
    Increase the number of observable strongly correlated DIBs. There are two avenues to explore. The first is to use spectral deconvolution to recognize and disentangle blended DIBs as well as other "spurious" spectral features. This method would be most effective when combined with spectral modeling in order to facilitate the separation of the various spectral features. An example of this is provided in Figure 9 for λ6196 where there are multiple small and narrow spectral features that do not appear to be associated with the primary carrier spectrum. The second avenue is to improve the S/N for the observations. This would directly impact the ${{f}_{\mathrm{EW}}}^{2}$ term in Equation (1) for which even a modest improvement in fEW, such as a factor of 1.5, would increase the number of observable strong correlations by a factor of ∼2.
  • 3.  
    Develop a semi-empirical or first principles model for the distribution of rotational states for a polar symmetric top molecule in the ISM. In this paper, we demonstrated the need for a two-temperature model to approximate the J, K state populations in order to support the common-carrier hypothesis for λ6614 and λ6196. The two-temperature approach was approximately implemented within the PGOPHER spectral model. A more first principles model would offer a more rigorous test of the common-carrier hypothesis. The first principles model should include (i) velocity-dependent state-to-state collisional excitation and de-excitation cross-sections between a candidate carrier and the most abundant ISM species, H, H+, H2, and He, (ii) predicted J, K populations based on the state-to-state collisional and radiative excitation and de-excitation rates for the local ISM physical conditions and interstellar radiation field, and (iii) simulated DIB absorption spectra based explicitly on the predicted ground state J, K populations. We note that the first principles model for the ground electronic state J, K populations would not obviate the need for the sharp rotational state cutoffs which we associate with the upper electronic state of the λ6614 DIB.

6. Summary and Conclusions

We used a combination of spectral deconvolution and spectral modeling to analyze observed spectra for the λ6614 and λ6196 DIBs. We determined physically reasonable, self-consistent spectroscopic parameters for both DIBs, indicating that they can originate from the same molecular carrier. Good spectral fits to both bands were demonstrated for two sightlines toward HD 145502 and HD 179406, representative of the range of spectral diversity typically observed for the DIBs.

Spectral deconvolution is a carrier-impartial technique for revealing finer spectral features than may be evident in the original data. Spectral deconvolution of λ6614 uncovered several key characteristics of its molecular carriers: (1) the central PQR structure arises from two comparably strong and slightly separated bands with similar shapes (i.e., a doublet), and (2) there is a red-tail residual whose shape is very different from the symmetric PQR components and that is likely due to a different carrier. The relative weights and the splitting of the doublet components were shown to be consistent with 13C isotope shifts at typical interstellar abundances.

We applied the PGOPHER spectral model to an analytic spectral profile derived from the deconvolution of a λ6614 DIB and derived physically plausible spectroscopic constants for its molecular carrier. A striking result is the sharp cutoff of the maximum values of the total rotational quantum numbers at a relatively low energy corresponding to 19.5 K. The sharp cutoffs were attributed to a shallow pretransformation barrier in the upper electronic state of the λ6614 DIB. Another key outcome is that the λ6614 spectrum could be transformed into one closely resembling that for λ6196 simply by varying the Coriolis constant, ζ. The spectral fit to λ6196 also required a lower rotational temperature, which was found to be consistent with a two-temperature model for a polar symmetric top molecule.

The spectral analysis supports two types of oblate, polar, symmetric top carriers: a compact, puckered PAH and a fullerane. The preferred PAH carrier is corannulene (C20H10), including its ionic and dehydrogenated symmetric top variants. However, we have not ruled out circumtridene (C36H12) and its variants. The favored fullerane carriers include C20H and C28H and their ionic forms.

We presented an analysis, based on the well-characterized distributions of DIB EWs and associated S/N levels, which implies there are numerous highly correlated DIB pairs. However, the number of observable correlations is severely limited due to a combination of S/N and DIB blending considerations. We suggest that spectral deconvolution can be used to remove the effects of DIB blending and potentially increase the number of identified highly correlated DIBs.

In conclusion, we have presented a detailed spectral analysis that establishes that the λ6614 and λ6196 DIBs could originate from the same molecule.

L.S.B. and R.M.S. appreciate funding for this project from Spectral Sciences, Inc., and L.S.B. appreciates additional funding from Maine Molecular Sciences. G.A.G. acknowledges the support of the Russian Science Foundation (project 14-50-00043, area of focus Exoplanets) for support of the experimental part of this work. The authors thank D. Lynch (Thule Scientific) and J. Gelbord (Spectral Sciences, Inc.) for a thorough review of the draft manuscript, and J. Quenneville (Spectral Sciences, Inc.) for technical discussions. The authors thank the anonymous reviewer whose insightful comments have resulted in significant improvements to this paper.

Appendix A: Sensitivity of the Spectral Model to Parameter Variations

In this appendix, we explore the sensitivity of the spectral model to the spectroscopic parameters when fitting the dominant PQR component in the λ6614 band of HD 145502. We demonstrate the impact of varying several different parameters in Figures 1521. For Figures 1519, one parameter is varied while the rest are fixed at the best-fitting values for λ6614, while in Figure 20 multiple parameters are allowed to vary simultaneously. Lastly, Figure 21 demonstrates the effect that variations in Trot have upon λ6196. We present parameter uncertainty estimates based on visually comparing overlays of the perturbed model spectrum to the spectral fit.

Figure 15.

Figure 15. Sensitivity of the spectral model fit to variations in Jmax for an analytic fit component (see Figure 6) of the deconvolved HD 145502 λ6614 data. The blue curve corresponds to the spectral fit to the analytic component. The spectral calculations in this section and throughout the paper maintain the constraint Jmax = Kmax discussed earlier.

Standard image High-resolution image
Figure 16.

Figure 16. Same as for Figure 15 except on an expanded frequency scale to capture the full extent of the modeled spectral profile for a value of Jmax (200) that is sufficiently large that it does exhibit a discernible cutoff for the P and R branches for Trot = 50 K.

Standard image High-resolution image
Figure 17.

Figure 17. Sensitivity of the spectral model fit to variations in ζ for the analytic fit component (see Figure 5) for the deconvolved HD 145502 λ6614 data. The blue curve with a substantial and negative ζ = −0.60 corresponds to the spectral model fit to the analytic component. In contrast, the model spectrum with a substantial and positive ζ = +0.5 more closely resembles the structure seen for λ6196.

Standard image High-resolution image
Figure 18.

Figure 18. Sensitivity of the spectral model fit to variations in Trot ${}_{}$for the analytical component (see Figure 5) for the deconvolved HD 145502 λ6614 data. The blue curve corresponds to the spectral fit to the analytic component.

Standard image High-resolution image
Figure 19.

Figure 19. Sensitivity of the spectral model fit to variations in B/C (only B was varied) for one of the identically shaped analytic components fitted to the deconvolved HD 145502 λ6614 data (see Figure 5). The blue curve corresponds to the spectral fit to the analytic component.

Standard image High-resolution image
Figure 20.

Figure 20. Comparison of spectral model fits for HD 145502 λ6614 for different assumed values of B/C in which all the other parameters were allowed to vary. The fit for an assumed compact PAH geometry, B/C = 2.0, is virtually identical to the fit presented earlier for B/C = 2.44 (see fit parameters in Table 1). Only two of the other parameters required an adjustment in order to obtain this fit, Jmax = 30 vs. 26 and ζ = −0.33 vs. −0.60 for B/C = 2.0 vs. 2.44.

Standard image High-resolution image
Figure 21.

Figure 21. Sensitivity of the spectral model fit to variations in Trot for the deconvolved HD 145502 λ6196 data (see Figure 9). The blue curve corresponds to the best model fit to the data presented earlier (see fit parameters in Table 1).

Standard image High-resolution image

Figures 15 and 16 shows that the widths of all of the branches and the peak locations of the P and R branches are sensitive to the assumed maximum value for the rotational quantum number cutoff, Jmax. The estimated spectral fit uncertainty for this parameter, assuming a sharp cutoff, is δJmax = ±1.

The truncation of the λ6614 spectrum reduces its observable EW to about 30% of its untruncated value. This implies that the "true" EW of λ6614 is the second largest of all DIBs, about 40% of that for λ4428 (Hobbs et al. 2008).

Figure 17 shows that the shape and width of the modeled spectrum are very sensitive to the assumed value of the Coriolis constant over its physically allowed range, −1 ≤ ζ ≤ 1. The PQR branch structure is seen for ζ < 0, whereas for ζ > 0 only P and R branches are evident (i.e., no Q branch), and these coalesce to form a single-peaked shape near ζ = −1. The width (FWHM) and separation of the P and R branch peaks for the modeled spectrum decreases substantially over the allowed range of ζ. The estimated uncertainty for this parameter is δζ = ±0.02.

Figure 18 reveals that the shape of the band is insensitive to Trot. For Trot > 25 K, the differences are hard to discern. For Trot = 10 K, the changes are apparent, but modest. The insensitivity to Trot is a consequence of the low rotational energy associated with the rotational quantum number cutoff (EJmax ∼ 13.5 cm−1 equivalent to 19.5 K). Because of the insensitivity of the shape to Trot, we can only establish an estimated lower limit of Trot > ∼25 K.

Figure 19 shows that the shape of the band is mildly sensitive to the value of a rotational constant. For these simulations, we fixed the value of C and let B vary. While one can discern a shape change for a change of ∼±5% in a rotational constant, it is misleading to consider the rotational constants as completely independent parameters. As discussed in Bernstein et al. (2015b), the shape of a band depends on the product of the rotational constants and temperature (e.g., BTrot), not on their individual values. In the current effort, we have shown that the shape also depends on Jmax and ζ. Hence, there is some ambiguity with regard to establishing a unique set of spectroscopic constants that can be associated with a specific molecule. As discussed earlier, other factors must be considered in order to break this ambiguity and constrain the selection of a preferred molecular carrier. In that regard, Figure 20 shows that a plausible fit can be found for a compact PAH geometry (B/C = 2.0), if one allows other parameters to vary.

The fit for λ6196 was constrained by the spectroscopic parameters determined for λ6614. We considered variations in ζ and Trot that were physically consistent with a common ground state but different upper transition states. In Figure 17, we saw that the shape and width for λ6196 could be mostly attributed to a change in ζ. The remaining width reduction follows from lowering Trot, as attributed earlier to the plausibility of a two-rotational temperature model for a polar symmetric top molecule. The sensitivity of the fit for λ6196 to Trot is shown in Figure 21. The estimated uncertainty is δTrot = ±0.5 K.

Appendix B: Spectroscopic Origins and Effects of Rotational Non-equilibrium

The more detailed spectroscopic explanation begins with the basic expression for the rotational energy of a symmetric top molecule (Herzberg 1989),

Equation (2)

where Bv is the vibrational-state-dependent rotational constant for the rotation axes perpendicular to the highest symmetry axis, Cv is the rotational constant for rotation about the highest symmetry axis, J is the total angular momentum quantum number, K is the quantum number for rotation about the highest symmetry axis, and ζv is the Coriolis constant. The ± notation relates to the relative rotational directions of the vibrational and rotational motions. For simplicity, in the following discussion we take v to refer to excitation in a single, specific mode (i.e., v = 0 is the ground vibrational state and v = 1 is the first vibrational excited state). We assume that the excited vibrational mode that gives rise to the DIB spectrum is doubly degenerate, i.e., its v = 1 state is comprised of two equal-energy levels. In this case, ζv = 0 for the ground and −1 ≤ ζ ≤ 1 for the first excited states. As demonstrated earlier (see Figure 17), a degenerate mode is required in order to produce the different shapes and widths seen for the λ6614 and λ6196 DIBs. For a nondegenerate vibrational mode, ζv = 0 for the ground and all excited vibrational states.

We focus on the role of the Cv term in Equation (2) in influencing the shape and width of a DIB band. This term isolates the contributions of transitions arising from the C rotational axis (see hot Trot axis in Figure 2), and it also includes the Coriolis interaction. The selection rules for a perpendicular electronic transition between a nondegenerate and a degenerate vibrational mode, as assumed here, are ΔJ = 0, ± 1 and ΔK = ±1 (Herzberg 1966). When these rules are applied to Equation (2) and we consider the dominant transition component for the Cv term, the result is

Equation (3)

This contribution occurs for all the bands and subbands of the P, Q, and R branches (Herzberg 1966). There are some smaller constant terms and higher-order K2 contributions, but their effect on the shape and width is minimal. In the limit of ζv = 1, ΔFv = 0. This means that only the transitions associated with cold rotational states (i.e., the Bv term in Equation (2)) contribute to the band shape and width. For λ6196, we found that ζv = 0.63 (i.e., in the vicinity of the ζv = 1 limit) and that very cold rotational temperatures, Trot = 3.5 and 7.0 K, were determined for the two sightlines. In contrast, for λ6614 we found that ζv = −0.60, which means the contribution of the Cv term is substantial, resulting in a much higher effective rotational temperature. For λ6614, we determined rotational temperatures of Trot = 18 and 50 K for the two sightlines.

Figure 22 provides a visualization of the role of the Coriolis constant in determining the sensitivities of a DIB profile to various spectroscopic parameters. For simplicity, we consider the limiting values of the Coriolis constant, ζ = ±1, because this provides the greatest contrast with respect to the differences of the spectral profiles and their sensitivities to various spectroscopic parameters. For ζ = +1, there is little sensitivity to large changes in the values of C and Jmax. In contrast, there is significant sensitivity to B and the cold Trot. For ζ = −1, there is significant sensitivity to all of the spectroscopic parameters.

Figure 22.

Figure 22. Sensitivity of the modeled spectrum at the limits of the Coriolis constant, ζ = 1 (left plot) and ζ = −1 (right plot), to variations in several spectroscopic parameters. A broadening parameter of γG = 0.30 cm−1 was used for all of the calculations. A reference curve is defined for each limit (black curve and black font) that resembles the dominant spectral component of the λ6196 (left plot) and λ6614 (right plot) spectra.

Standard image High-resolution image

Appendix C: Populations of the (J, K) States for a Symmetric Top Molecule

Several key findings of this effort, namely the origin of the sharp cutoff of the rotational quantum numbers, Jmax = Kmax, and the applicability of the two-rotational-temperature model, are supported by quantitative consideration of the physical processes controlling the (J, K) state populations of the symmetric top carrier. The balance between collisional excitation and relaxation and radiative excitation and decay, in the ISM, determines the (J, K) state populations. The population analysis presented below borrows from that for the RTE model as described in detail by Loren & Mundy (1984) and applied to the ISM microwave spectrum of the prolate symmetric top molecule CH3CN.

We assume that the rates for collisional excitation and de-excitation of a rotational state are comparable to the gas kinetic collision rate between the carrier and the ambient species. The gas kinetic collision rate constant is estimated from

Equation (4)

where σ is the collision diameter and vH is the average velocity of a H atom. The factor of 1/2 accounts for the average projection of the oblate top area in the collision direction. For a H atom colliding with corannulene, we estimate a collision diameter of σ = 4.5 Å and an average velocity of vH = 1.0 × 105 cm s−1 for Tgk = 50 K. The estimated gas kinetic rate constant is kgk = 3.3 × 10−10 cm3 s−1. The rate constants for the other abundant CNM species, He and H2, are essentially the same as for H; their collision diameters are slightly larger but smaller collision velocities offset this. Using a typical average density for a molecular cloud (Goldsmith 2013) of n = 75 cm−3, the collision rate is 75kgk = 2.5 × 10−8 s−1. Below, we compare the collision rate to the radiative decay rates for the tumbling and symmetric spinning motions in order to estimate the effect of rotational temperatures for these motions.

We first consider the tumbling motion and show that (1) the J distribution for a K stack corresponds to a cold effective rotational temperature, and (2) Jmax = Kmax is a good approximation. The radiative decay rate for a (J, K) state for the tumbling motion, where K is fixed and J > K, is given by its Einstein A coefficient for J, K ≫ 1 (Ali-Haïmoud 2014),

Equation (5)

where h is Planck's constant, μ is the dipole moment, which is aligned with the highest-order symmetry axis, B is the rotational constant for rotation perpendicular to the symmetry axis (i.e., the end-over-end tumbling motion), and J ≥ 1 is the quantum number for the upper state of the transition. For a symmetric top molecule, the selection rules for rotational emission are J → J–1 and ΔK = 0. Hence, a symmetric top can reduce the rate of its tumbling motion through radiative decay, whereas the spin rate about its highest-order symmetry axis remains constant because radiative decay for this motion is forbidden.

Evaluation of ${A}_{J,K\to J\mbox{--}1,K}$ in Equation (5), using the carrier parameter values μ = 2D and B = 0.02 cm−1, yields

Equation (6)

Comparison of the radiative decay rates to the ISM collision rate leads to a constraint on the "excess" angular momentum (i.e., Jmax > Kmax) available for the tumbling motion. We seek the value of J for a given K for which the radiative decay rate is approximately equal to the collision rate. For higher values of J, the radiative decay rate significantly exceeds the collision rate resulting in a sharp falloff in the population for the higher J states. The radiative decay rate for J = Kmax + 1 and K = Kmax = 26, for which ${A}_{J,J\mbox{--}1}=5.8\times {10}^{-8}$ s−1, exceeds the collision rate, 2.5 × 10−8 s−1, but about a factor of two. This means that the J = Kmax + 1 state will most likely emit a photon, emission probability = (1 – e−2) = 0.86, before it experiences a collision. This finding supports the Jmax = Kmax assumption, since even one additional excitation of J will most often radiatively decay before the next excitation collision can occur.

However, for higher cloud densities, the collision rate is fast enough to support several J states above Kmax. At the upper density boundary for the DIB-containing clouds, the density is n ∼ 500 cm−3 (Snow & McCall 2006) and Jmax = Kmax + 6. This increase in Jmax will cause a noticeable outward extension of the P and R branch peaks and edges by about 6(2B) = 0.24 cm−1. On the other hand, the modeled λ6196 spectra will not be significantly impacted by an increase in Jmax because their much lower Trot render the spectra relatively insensitive to the higher J and K states.

As just discussed, the K stacks for the higher K values support just a few J levels. However, the lower K values can support considerably more J levels. For the lower energy states, one must also include the effects of radiative excitation and de-excitation driven by the ambient radiation field. For K = 0, the problem of determining the populations of the J levels is analogous to that for a diatomic molecule. We base our analysis on the analytic model of Oka et al. (2013), which predicts the J distribution for a linear molecule subject to the combined effects of collisional and radiative excitation and relaxation. Bernstein et al. (2015b; see Figure 5) applied Oka's model over a wide range of molecular and environmental conditions that encompass the problem under consideration in this work. The key result is that Trot spans a narrow range of ∼3–8 K assuming a range of collision rates of 10−8–10−7 s−1, a range of rotational constants of 0.01–0.02 cm−1, μ = 2 D, Tgk = 100 K, and the CMB radiation field. As discussed in Appendix B, the λ6196 spectrum arises primarily from the tumbling motion transitions; hence, it should be characterized by a cold Trot. The retrieved Trot of 3.5 and 7.0 K for the two λ6196 spectra are consistent with the ∼3–8 K temperature range deduced from the theoretical analysis.

In contrast to the tumbling motion, the spinning top motion cannot emit via dipole-allowed transitions. Ignoring forbidden transitions, this means that the effective rotational temperature for the K stacks should be equal to the gas temperature. Snow & McCall (2006) suggest a range of gas temperatures of ∼30–100 K for the DIB carrier environments. As discussed in Appendix B, the λ6614 spectrum is strongly dependent on spinning top transitions (i.e., ΔK = ±1); hence, it should be characterized by a Trot comparable to the gas temperature. The retrieved Trot of 18 and 50 K for the two λ6614 spectra are consistent with the estimated ∼30–100 K gas temperature range.

We consider the potential for radiative relaxation of the K distribution through so-called forbidden radiative transitions. These include the centrifugal distortion and quadrupole-allowed transitions for which ΔK = 3 and 2, respectively. We estimate the order of magnitude of their Einstein A coefficients for our prototype carrier by approximately scaling from known transition rates using

Equation (7)

where A0 is the Einstein A coefficient for a known transition, ν is the transition frequency, and m = 6 or 5 for centrifugal distortion and quadrupole-allowed transitions, respectively. For the centrifugal distortion of ${{{\rm{H}}}_{3}}^{+}$ (Pan & Oka 1986), A0 < 0.01 s−1 for ν0 ∼ 200 cm−1 (i.e., upper limit estimate), and for the prototype carrier, ν = 6BJmax = 6*0.02*26 = 3.1 cm−1, therefore A = 0.01(3.1/200)6 = 1.4 × 10−13 s−1. This radiative decay rate is much smaller than the gas kinetic collision rate, discussed above, of 2.5 × 10−8 s−1. Therefore, radiative relaxation via centrifugal distortion allowed transitions is not a significant mechanism for the relaxation of the K distribution. The same outcome applies to quadrupole-allowed transitions, where A = 4.8 × 10−10(2.1/587)5 = 2.8 × 10−22 s−1 based on the scaling of the S(0) transition for H2 (Wolniewicz et al. 1998).

Appendix D: Physical Origin for the Sharp Rotational Level Cutoff for λ6614

We conjecture that the essential ingredients for a carrier to exhibit a sharp cutoff include (1) a low "pretransformation" barrier in the upper electronic state, and (2) an effectively irreversible transformation with line widths larger than the width of the DIB spectrum. For polyatomic molecules, transformations other than dissociation can occur, such as a geometry change or the migration of a H atom to a different C atom site. In order to observe a cutoff, the barrier height must be commensurate with the available vibrational and rotational excitation in the excited electronic state. The available rotational energy for the DIB carrier is small, on the order of ${E}_{\mathrm{rot}}\sim {{{CK}}^{2}}_{\max }=0.01* {26}^{2}=7$ cm−1. The available vibrational energy corresponds to a quantum of excitation in a degenerate vibrational mode of the upper electronic state, as necessitated by the need for Coriolis rotation–vibration coupling. For a PAH or fullerane carrier, the vibrational excitation requirement constrains the barrier height to a vibrational energy range of ∼100–3000 cm−1. Thus, the vibrational excitation supplies nearly all of the energy required to surmount the barrier. However, the rotational energy is needed to bridge the small gap between the vibrational excitation and the barrier peak and to facilitate the transformation via a rotation–vibration coupling mechanism. For the proposed carriers, the Coriolis coupling between an in-plane degenerate vibration and the spinning top rotational motion could satisfy the coupling requirement.

The second ingredient involves a fast and essentially irreversible transformation. The required speed of the transformation can be estimated using the Heisenberg uncertainty principal, which ties the line width of a transition, Γ(cm−1), to the lifetime of its upper transition state, δt (s),

Equation (8)

where c (cm s−1) is the speed of light. In order for a transition to blend in with the baseline, its line width needs to be several times larger than the bandwidth, which implies Γ  ∼  4 cm−1 for λ6614 (see Figure 7). The resultant lifetime is 1.3 ps. For comparison, a representative vibrational energy of 1000 cm−1 corresponds to a vibrational period of 0.03 ps. Thus, the transformation can occur on a timescale corresponding to roughly 10–100 vibrational periods.

The irreversibility constraint is satisfied for a diatomic molecule via rapid dissociation, once the barrier is surmounted. For a polyatomic molecule, irreversibility could be satisfied by fast intramolecular vibrational energy redistribution (IVR) in the transformed molecule. In this picture, the initially transformed state would be rapidly subsumed into an ocean of iso-energetic states, resulting in a miniscule probability of reversing the transformation process. It is widely held that the DIBs and the UIR are both due to PAHs. The widths of the UIR bands, of order 10 cm−1, are ascribed to IVR (Boulanger et al. 1998). Thus, it seems feasible that λ6614 could exhibit a comparable width in a transformed molecular state.

The possibility of a carrier meeting the requirements for a sharp cutoff is bolstered by the existence of a similarly complex, aromatic molecule that demonstrates most of the requirements. We are referring to the well-studied trans to cis isomerization of stilbene, C14H12 (Baskin et al. 1996; Quenneville & Martınez 2003). Stilbene's trans singlet ground state, S0, has a strongly allowed absorption to its lowest singlet excited state, S1. There is a small barrier, ∼1200 cm−1, separating the trans and cis isomers on the excited S1 surface. Analogous to the predissociation barrier for OH, the stilbene barrier also arises from an avoided crossing. The lifetimes of states near the barrier are of order ∼1 ps, and once the cis transformation is complete, rapid IVR occurs. The coupling, if any, of the rotational and vibrational motions near the top of the barrier has yet to be theoretically or experimentally verified.

How does the stilbene analogy relate to the transition for the λ6614 carrier? We speculate that the most likely parallel is a geometry change in the lowest excited state, such as transforming from a puckered to a flat disk for corannulene, migration of a H atom, and warping to an asymmetric shape.

Please wait… references are loading.
10.3847/1538-4357/aabd85