Chapter 3
The Grain Properties of Nearby Galaxies

 3.1 Spectral Energy Distribution Modeling
  3.1.1 Radiative Transfer
   3.1.1.1 Definition of the Main Radiative Transfer Quantities
   3.1.1.2 The Radiative Transfer Equation
   3.1.1.3 Approximations for Clumpy Media
   3.1.1.4 Rigorous Solutions
  3.1.2 Approximate Treatments of the Mixing of Physical Conditions
   3.1.2.1 The Historical Model: the MBB
   3.1.2.2 A Phenomenological, Composite Approach
   3.1.2.3 Panchromatic Empirical SED Models
   3.1.2.4 The Matryoshka Effect
  3.1.3 Application to Nearby Galaxies
   3.1.3.1 The Different Types of Galaxies
   3.1.3.2 Large-Scale Dust Distribution in Galaxies
   3.1.3.3 Constraining the Grain Opacity
   3.1.3.4 Constraining the Size Distribution
 3.2 Studies Focussing on Specific Spectral Domains
  3.2.1 Scrutinizing Mid-IR Spectra
   3.2.1.1 The Aromatic Feature Spectrum
   3.2.1.2 Spectral Decomposition Methods
   3.2.1.3 PAH Band Ratio Studies
   3.2.1.4 Variations of the Aromatic Feature Strength
  3.2.2 Long-Wavelength Properties
   3.2.2.1 The Elusive Submillimeter Excess
   3.2.2.2 The Anomalous Microwave Emission
 3.3 Dust in Relation with the Gaseous and Stellar Contents
  3.3.1 The Phases of the ISM
   3.3.1.1 The Neutral Atomic Gas
   3.3.1.2 The Ionized Gas
   3.3.1.3 The Molecular Gas
  3.3.2 Dust as a Diagnostic Tool
   3.3.2.1 Dust to Study Star Formation
   3.3.2.2 Photodissociation Regions
   3.3.2.3 The Molecular Gas and its Dark Layer

Without data, you’re just another person with an opinion.
 
(Attributed to W. Edwards DEMING)

In this chapter, we review how the models presented in Sect. 2.3 are used to derive the grain properties of nearby galaxies. The term “dust properties” is vague. In the literature, it often indistinctively encompasses the three following categories (Galliano, Galametz, & Jones, 2018).

The dust mixture constitution
is characterized by:
The dust physical conditions
are the state a grain enters, when exposed to a particular environment:
The dust observables
arise from a grain mixture experiencing a particular set of physical conditions.
Emission
of partially polarized components (Sect. 2.2.2.2): (i) a thermal continuum (IR to mm; Sect. 2.2.2.1); (ii) molecular and solid-state features (MIR; Sect. 2.2.2.1); (iii) a possible microwave emission (cm; Sect. 2.2.2.3); (iv) a possible luminescence (visible; Sect. 2.2.2.3).
Absorption
of the light from a background source by (Sect. 2.2.1): (i) a continuum (X-ray to MIR; Sect. 2.2.1.1); (ii) atomic, molecular and solid-state features (X-rays, UV and MIR; Sect. 2.2.1.2), including DIBs (visible; Sect. 2.2.1.5) and ices (MIR; Sect. 2.2.1.2). The escaping light can be partially polarized as a result (Sect. 2.2.1.4).
Scattering
of the light from a bright source in our direction, and its polarization (X-rays to NIR; Sect. 2.2.1).
Depletion
patterns seen through the gas-phase elemental abundances (Sect. 2.2.3).

3.1 Spectral Energy Distribution Modeling

SED modeling is one of the main methods to empirically derive the dust properties of a region or a galaxy (cf. e.g. Galliano et al., 2018, for a review). The inherent complexity of astrophysical sources requires to account for the diversity of physical conditions within the studied region. The treatment of radiative transfer, even in an extremely approximated fashion, is thus necessary.

3.1.1 Radiative Transfer

Radiative transfer is the method solving the propagation of multiple rays of light, emitted by one or several sources, through a macroscopic heterogeneous medium. It accounts for the scattering, absorption and emission, at each point and along each direction, in the studied region.

3.1.1.1 Definition of the Main Radiative Transfer Quantities

Radiative transfer deals with several quantities that are often improperly defined or mixed together in the literature: intensity, flux, emissivity, brightness, etc. We have already seen some of these quantities in Sect. 1.2.4. We now define them and explicit their differences (cf. Chap. 1 of Rybicky & Lightman, 1979, for a complete review). In what follows, we assume stationary systems. The time variable, t, is used only to denote constant rates.

The moments of the specific intensity. The primary radiative transfer quantity is the specific intensity or brightness (cf. Fig. 3.1.a):

Iν(ν,r,𝜃,ϕ) dE dtdAdΩdν. (3.1)

The specific intensity is the electromagnetic energy, E, per unit time, t, area, A, solid angle, Ω, and frequency, ν. It therefore quantifies the infinitesimal power carried by a monochromatic light ray. This quantity depends on the position in the region,r, and on the direction of propagation, (𝜃,ϕ). We adopt the spherical coordinate conventions used in physics (Fig. 3.1; cf. Appendix C.1):

The first moments of the specific intensity, relative to its angular distribution, are physically meaningful.

The mean intensity
is the zeroth order moment of Eq. (3.1):
Jν(ν,r) 1 4π sphereIν(ν,r,𝜃,ϕ)dΩ. (3.3)

It is the specific intensity averaged over all directions. It is often used to quantify the ISRF, 4πJν accounting for rays coming from all directions. In the isotropic case, we have Iν(ν,r,𝜃,ϕ) = Jν(ν,r), (𝜃,ϕ).

The net flux
(cf. Fig. 3.1.b) is the first order moment of Eq. (3.1):
Fν(ν,r) sphereIν(ν,r,𝜃,ϕ) cos 𝜃dΩ. (3.4)

It represents the monochromatic power per unit area passing through a surface element perpendicular to . The cos 𝜃 factor is there to account for the reduction of the density of rays that are not perpendicular to the surface. In the isotropic case, Fν = Jν2π11 cos 𝜃d cos 𝜃 = 0, because there is the same amount of flux passing through the area in both directions.

The radiation pressure
(cf. Fig. 3.1.c) is the second order moment of Eq. (3.1):
pν(ν,r) 1 c sphereIν(ν,r,𝜃,ϕ) cos 2𝜃dΩ (3.5)

It is the momentum flux carried by the photons, as p = hνc is the momentum of a single photon. The cos 2𝜃 term comes from two sources: (i) one cos 𝜃 comes from the reduced fraction of inclined rays, similar to the flux; (ii) the second cos 𝜃 factor comes from the fact the pressure is the momentum vector component that is perpendicular to the surface.

PIC
Figure 3.1: Moments of the specific intensity. Panel (a) represents the specific intensity (Eq. 3.1). It is the monochromatic power of light rays per area, dA, and within the solid angle dΩ. Panel (b) demonstrates the calculation of the flux (Eq. 3.4). We have represented several rays, with different directions and intensities. The effective area perpendicular to rays going through dA, but that are inclined at an angle 𝜃, is only cos 𝜃dA. Panel (c) represents the radiation pressure (Eq. 3.5). It is similar to panel (b), except that we have shown the component of the momentum vector, p, perpendicular to the surface, |p| cos 𝜃. Only this component contributes to the pressure on dA. Panel (d) represents the energy density (Eq. 3.6). Between times t and t + dt, photons going through dA are encompassed within the cylinder of volume cdtdA. Licensed under CC BY-SA 4.0.

Specific energy density. The radiative energy within a volume element at a given time is the specific energy density:

uν(ν,r,𝜃,ϕ) dE dV dΩdν = dE cdtdAdΩdν. (3.6)

The second equality comes from the fact that dV = cdtdA, dV being a volume element (cf. Fig. 3.1.d). The specific intensity is a power per unit area, whereas the energy density is an energy per unit volume. Both quantities are linked, combining Eq. (3.1) and Eq. (3.6):

uν(ν,r,𝜃,ϕ) = Iν(ν,r,𝜃,ϕ) c . (3.7)

If we integrate Eq. (3.6) over all directions, we get, using Eq. (3.3):

Uν(ν,r) = sphereuν(ν,r,𝜃,ϕ)dΩ = 4π c Jν(ν,r). (3.8)

Emission coefficient and emissivity. The monochromatic emission coefficient is the power radiated in a given direction, per unit volume and frequency:

jν(ν,r,𝜃,ϕ) dEem dtdV dΩdν. (3.9)

We have also seen, in Sect. 1.2.4, the emissivity:

𝜖ν(ν,r,𝜃,ϕ) 4π dEem dtdmdΩdν = 4π dEem dtρdV dΩdν, (3.10)

where ρ is the mass density of the ISM, and dm = ρdV , its mass element. The factor 4π, in Eq. (3.10), makes 𝜖ν the solid angle fraction of monochromatic power emitted in a given direction, per unit mass. In Sect. 1.2.4, we were considering the volume and mass of a grain, whereas, here, we are considering the volume and mass of the ISM. Combining Eq. (3.9) and Eq. (3.10), we get:

jν(ν,r,𝜃,ϕ) = ρ(r) 4π 𝜖ν(ν,r,𝜃,ϕ). (3.11)

Extinction coefficient and opacity. The amount of specific intensity absorbed and scattered along an infinitesimal path length, dl, in the direction (𝜃,ϕ), is the extinction coefficient, α 0, defined such that:

dIν(ν,r,𝜃,ϕ) dl = α(ν,r)Iν(ν,r,𝜃,ϕ). (3.12)

Similarly to the convention we have adopted for κ in Sect. 1.2.4, we pose α αext = αabs + αsca to distinguish absorption and scattering.. Eq. (3.12) can be expressed with microscopic quantities, assuming the absorbers and scatterers have a cross-section Cext(ν,r) and a density n(r):

α(ν,r) = n(r)Cext(ν,r). (3.13)

If the composition of the ISM is homogeneous, then Cext(ν,r) = Cext(ν). The opacity, that we have seen in Sect. 1.2.4, is related to α by:

α(ν,r) = ρ(r)κ(ν,r). (3.14)

Mean free path. The mean free path of a photon with frequency, ν, at position r can be defined as:

lmean(ν,r) 1 α(ν,r) = 1 ρ(r)κ(ν,r). (3.15)

It is the average length a photon will be able to travel before being absorbed or scattered. We can make the same remark as for the emissivity: in Sect. 1.2.4, we were considering the cross-section of dust particles per mass of grain, whereas Eq. (3.14) gives the cross-section of the whole ISM per mass of ISM. Table 3.1 gives typical values of lmean in a homogeneous medium, assuming the dust constitution of the THEMIS model (cf. Sect. 2.3.2.2).

 In the diffuse ISM (WNM; nH 0.3cm3; Table 3.6), the mean free path of a photon in the visible range is of the order of a kiloparsec.







HIM WNM CNM
Molecular clouds

nH = 0.003cm3 n H = 0.3cm3 n H = 30cm3 n H = 104cm3 n H = 106cm3






lmean(U)

139 kpc 1.39 kpc 13.9 pc 0.0417 pc 86.1 a.u.






lmean(B)

177 kpc 1.77 kpc 17.7 pc 0.0532 pc 110 a.u.






lmean(V )

223 kpc 2.23 kpc 22.3 pc 0.0669 pc 138 a.u.






lmean(R)

275 kpc 2.75 kpc 27.5 pc 0.0824 pc 170 a.u.






lmean(I)

358 pc 3.58 kpc 35.8 pc 0.107 pc 222 a.u.






lmean(J)

691 kpc 6.91 kpc 69.1 pc 0.207 pc 427 a.u.






lmean(H)

1021 kpc 10.2 kpc 102 pc 0.306 pc 632 a.u.






lmean(K)

1734 kpc 17.3 kpc 173 pc 0.52 pc 1073 a.u.






Table 3.1: Mean free path as a function of wavelength and density. These quantities were computed by rewriting Eq. (3.15) as 1lmean(λ) = nHmHY dustκ(λ), taking Y dust MdustMH from Table 2.4 and κ values from Table 2.5. We quote lmean at the same photometric bands as in Table 2.5. These values correspond to Solar metallicity, Z = Z. At first order, for metallicities Z 0.2Z, we can assume linearity: lmean(Z) lmean(Z) × ZZ (cf. Chap. 4). The different ISM phases quoted here (HIM, WNM, CNM) will be defined in detail in Sect. 3.3.1.
3.1.1.2 The Radiative Transfer Equation

The radiative transfer equation accounts for the variation of the specific intensity under the effects of absorption, scattering and emission (cf. Steinacker et al., 2013, for a review in the case of a dusty medium). It is schematically represented in Fig. 3.2. This equation can be written:

dIν(ν,r,𝜃,ϕ) dl = αabs(ν,r)Iν(ν,r,𝜃,ϕ) absorption αsca(ν,r)Iν(ν,r,𝜃,ϕ) scattering out of the sightline + αsca(ν,r)2π11Φ(cos 𝜃,ν)I ν(ν,r,𝜃(𝜃),ϕ(𝜃))d cos 𝜃 scattering in the sightline + jνdust(ν,r) dust emission + jν(ν,r) stellar emission. (3.16)
Scattering in the sightline:
this term is the integral of αscaIν (i.e. the scattered intensity) over the phase function, Φ (Eq. 1.40). This expression depends on the relative angle, 𝜃, between the incident rays and the scattered direction (𝜃,ϕ). This is the term that makes this equation an integro-differential equation, coupling all the directions together. This is why, numerical methods are required to solve Eq. (3.16). If we assume isotropic scattering (i.e. cos 𝜃 = 0, corresponding to the Rayleigh regime; cf. Sect. 1.2.2.3), this term simplifies and becomes: αsca(ν,r)Jν(ν,r).
Emission:
we have assumed that the dust and stellar emissions are both isotropic, (i.e. independent of 𝜃 and ϕ). This is a reasonable assumption in the ISM. If we also assume that there is only one grain species, and that this species is at equilibrium with the radiation field, we can explicit: jνdust(ν) = α(ν)B ν(ν,Teq). This simplification still contains the problem that to determine the equilibrium temperature, Teq, we need to integrate αabsIν over all directions, and all frequencies.

Eq. (3.16) can be rewritten, using the optical depth, τ, defined such that dτ(ν) = α(ν,l)dl, or:

τ(ν,l) =0lρ(l)κ(ν,l)dl. (3.17)

Eq. (3.15) implies that τ(lmean) = 1. At a given wavelength, a medium is: (i) optically thin or transparent, if τ « 1; and (ii) optically thick or opaque, if τ » 1. Replacing dl by dτα in Eq. (3.16), we obtain:

dIν(ν,r,𝜃,ϕ) dτ = Iν(ν,r,𝜃,ϕ) + Sν(ν,r,𝜃,ϕ), (3.18)

where the source function, Sν, includes all the terms added to the specific intensity:

Sν(ν,r,𝜃,ϕ) = ω̃(ν,r)2π11Φ(cos 𝜃,ν)I ν(ν,r,𝜃(𝜃),ϕ(𝜃))d cos 𝜃+jνdust(ν,r) + j ν(ν,r) α(ν,r) . (3.19)

We will discuss exact numerical solutions to this equation in Sect. 3.1.1.4. For now, we discuss trivial solutions, when some processes are assumed negligible.

PIC
Figure 3.2: The radiative transfer equation. This figure represents only the different processes contributing to the variation of the specific intensity at one location, r, in one direction, (𝜃,ϕ). The blue sphere represents a volume element. The red spheres within represent dust grain and the cyan star represents an actual star that would be present in the volume element. Licensed under CC BY-SA 4.0.

Propagation in vacuum. Let’s assume we have a star of radius, R, and surface temperature, T, located at r = 0. The flux at r = R has to be integrated only over the hemisphere where a surface element of the star emits: Fν(R) = 2π01B ν(T) cos 𝜃d cos 𝜃 = πBν(T). If there is no ISM around the star, Eq. (3.16) simply becomes dIνdl = 0. The solution is thus Iν = Bν(T) along the directions coming from the star and 0 in all other directions. At an arbitrary distance, r, from the star, the solid angle it occupies is Ω = R2r2. The flux is thus Fν(r) = πBν(T)R2r2, which is the classic 1r2 dilution of the flux.

Emission only. Let’s assume we are observing, at submm wavelengths, a molecular cloud constituted of equilibrium grains at T 10 K, with opacity κ. At these wavelengths, the extinction is negligible. Eq. (3.16) is therefore simply dIνdl = ρκBν(T) within the cloud. The solution is thus Iν(l) = κBν(T)l0lρ(l)dl, where l0 is the position of the edge of the cloud along the direction of the sightline. If the density is constant, we get Iν(l) = ρκBν(T) × (l l0), which can be simplified as Iν(l) = τ(l)Bν(T). This expression of the brightness is often used in radio-astronomy.

Absorption only. Let’s assume we are observing a background star through a cold molecular cloud, in the MIR. At these wavelengths, the albedo is close to 0 (Sect. 1.2.2.3). If we make the assumption that the background star is much brighter than the thermal emission of the cloud, Eq. (3.16) simply becomes dIνdl = αIν. The solution is therefore Iν(l) = Iν exp τ(l) along the direction coming from the star and 0 in all other directions.

Emission and absorption. We can merge the two previous cases. It could correspond to a hot molecular cloud, observed at MIR wavelengths. Its thermal emission is absorbed by the cloud itself. The solution of Eq. (3.18), in this case, is:

Iν(τ) = Iν exp τ +0τ exp τ τS ν(τ)dτ. (3.20)

If the cloud contains a single grain species at temperature T, the solution becomes:

Iν(τ) = Bν(T) + Iν B ν(T) × exp (τ). (3.21)

If we look in a direction away from the background star, we get the classical self-absorption formula:

Iν(τ) = Bν(T) × 1 exp (τ) . (3.22)

This solution is displayed in Fig. 3.3. Eq. (3.22) has the following two limit regimes.

Optically thin:
if τ « 1, Iν τBν(T), which is the “emission only” solution. The emission from the cloud is a grey body (Sect. 1.2.4).
Optically thick:
if τ » 1, Iν Bν(T), which means the cloud is not anymore a grey body, but a perfect one.
PIC
Figure 3.3: Solution to the radiative transfer equation for an isothermal cloud, without scattering. We have assumed that we are looking at a background star (T = 6000 K), through an homogeneous cloud of grains at equilibrium temperature, T = 100 K, with the THEMIS opacity. We have diluted the stellar emission by a factor 106, which could correspond to a situation where the actual angular area of the star is 106 times the beam of the telescope. We show the SED according to Eq. (3.22), varying the optical depth in the V band, τV. We have included very large τV, up to 105, in order to demonstrate the asymptotic behavior of the dust emission tending toward a perfect black body. In typical ISM studies, it is however rare to find τV values higher than 100. Licensed under CC BY-SA 4.0.

Scattering and absorption with central illumination. Let’s assume that we have a homogeneous spherical cloud, of radius Rcl, and a central isotropically illuminating source, with monochromatic luminosity, Lν(λ). If Lνesc(λ) is the monochromatic luminosity escaping the cloud, the escape fraction can be defined as:

Pνesc(λ) Lνesc(λ) Lν(λ) = exp τeff(λ) . (3.23)

The second equality defines τeff as the effective optical depth of the medium.

Optically thick:
In the case of pure isotropic scattering (g = cos 𝜃 0), it can be shown that the net displacement of a photon after Nsca interactions is: leff = Nsca lmean (cf. Chap. 1 of Rybicky & Lightman, 1979). The probability a photon will be absorbed, at the end of a free path, is 1 ω̃. The mean number of free paths can thus be estimated by Nsca(1 ω̃) = 1, or: Nsca = 1(1 ω̃). The optical depth of the cloud is τ Rcllmean (Eq. 3.15 and Eq. 3.17). The same way, the effective optical depth can be written τeff Rclleff. We thus get:
τeff 1 ω̃τ. (3.24)
Optically thin:
in this case, there is a low probability of interaction. Eq. (3.23) tells us that τeff accounts only for photons that have been absorbed. We therefore simply have (cf. Chap. 1 of Rybicky & Lightman, 1979):
τeff = (1 ω̃)τ. (3.25)

This formula simply subtracts among the few photons that may have interacted with the grains those which have been scattered. Contrary to Eq. (3.24), it is valid for any value of g.

Városi & Dwek (1999, hereafter VD99) have proposed an empirical approximation to interpolate the two regimes of Eq. (3.24) and Eq. (3.25), resulting in the following escape fraction:

Pesccen(λ) exp (1 ω̃(λ))χ(λ)τ(λ) , (3.26)

with:

χ(λ) 1 1 2 1 exp τ(λ) 2 1 g(λ). (3.27)

They have benchmarked this approximation with a Monte-Carlo radiative transfer model (cf. Sect. 3.1.1.4). Both are in very good agreement in most of the astrophysically relevant parameter space. This solution is displayed in Fig. 3.4.a.

PIC PIC
Figure 3.4: Escaping radiation from a spherical cloud. In the three panels, we show the escaping SED of a homogeneous spherical cloud, made of THEMIS grains. We vary the V-band optical depth, τV. The illuminating source is a star (T = 3 × 104 K), with a bolometric luminosity L = 104L . In panel (a), we show the case of central illumination. All the power is in the central source. Panel (b) shows the case of uniform illumination. The total power of the sources within the cloud is L. Panel (c) shows external illumination. In this case, the flux at the surface of the cloud is L4πRcl2. Licensed under CC BY-SA 4.0.

Scattering and absorption with uniform illumination. VD99 have also proposed an approximation in the case where the internal illumination of the cloud is uniform. They start from the escape probability of a homogeneous sphere with uniform illumination, without scattering, given by Osterbrock (1989):

Pescnosca = 3 4τ 1 1 2τ2 + 1 τ + 1 2τ2 exp 2τ. (3.28)

A demonstration of this formula is given in Appendix C of VD99. VD99 find, with a recursive argument, that the following expression is in relatively good agreement with a Monte-Carlo radiative transfer model:

Pescuni(λ) Pescnosca(λ) 1 ω̃(λ) 1 Pescnosca(λ) . (3.29)

VD99 show that Eq. (3.29) perfectly agrees with the exact solution for a particular value of g(τ). The largest discrepancies, of about 20%, are found at high optical depth, for g close to 0. Eq. (3.29) is demonstrated in Fig. 3.4.b. We can see that, at a given τV, this geometry has an overall larger escape fraction than central illumination.

Scattering and absorption with external illumination. There are also expressions in the case of a homogeneous spherical cloud, externally illuminated by an isotropic radiation field. VD99 derive the following approximation for the absorbed fraction:

Pabsext(λ) 1 P escext(λ) 4τ(λ) 1 ω̃(λ) 3 Pescuni(λ). (3.30)

The absorbed fraction is indeed more relevant in the case of external illumination, as it would be difficult to observationally separate the escaping radiation from the cloud and the ambient ISRF. On the contrary the absorbed fraction is meaningful if one wants to evaluate the heating of the cloud. This approximation is demonstrated in Fig. 3.4.c.

The three formulae for spherical clouds, given in Eq. (3.26), Eq. (3.29) and Eq. (3.30), do not allow us to model the internal heating of the cloud. These expressions are indeed global escape and absorbed fractions, but they do not account for the gradient of illumination within the cloud that would lead to a gradient of heating rate.

3.1.1.3 Approximations for Clumpy Media

The ISM is a highly heterogeneous medium, with contrast densities of several orders of magnitude. A useful approximation is the clumpy medium, composed of: (i) a diffuse, uniform InterClump Medium (ICM), characterized by its density, nICM; and (ii) dense, spherical clumps, with density, nC, radius, rC, and volume filling factor, fV.

Effective optical depth of a clumpy medium. Let’s assume that we are looking at a background star through a cloud, and that the albedo of the grains is negligible. Along a given sightline, we have seen in Sect. 3.1.1.2 that the brightness in the direction of the star is Iν = Iν × exp τ. If the clumpy structure of the cloud is unresolved, the brightness measured in the telescope beam can be written as the sum of N sightlines, some passing through the ICM, others through the clumps:

Iνclumpy = Iν N i=1N exp τ i . (3.31)

This is the general expression. In the case of an homogeneous medium, Eq. (3.31) simplifies: Iν = Iν exp (τ hom), where τhom is the optical depth of the homogeneous medium. In order to have the same dust mass and opacity as in the clumpy medium, we need to have:

τhom 1 N i=1Nτ i. (3.32)

In the homogeneous medium, ρ × L is indeed the mass surface density of a cloud of depth L. If there is a statistical distribution of clumps in the beam, it must be identical in the clumpy medium. Thus, the brightness of the homogeneous medium is:

Iνhom = exp 1 N i=1Nτ i = i=1N exp (τ i) 1N. (3.33)

Invoking the arithmetic-mean/geometric-mean inequality 1 (e.g. page 456 of Cauchy, 1821), we conclude that: Iνclumpy I νhom. An important consequence of this result is that, from an observational point of view, we can miss a large mass of dust hidden in clumps. VD99 discuss this result in more detail.

 The effective optical depth of a clumpy medium is always lower than that of a homogeneous medium with the same dust constitution and dust mass.

The mega-grains approximation. Neufeld (1991) proposed a simple approach to explain the leakage of Ly-α photons by galaxies. He treated dusty gas clumps, in an empty ICM, as large grains with their own albedo and asymmetry parameter. This idea was then further developed by Hobson & Padman (1993), in cases where the ICM is non empty. They named it the mega-grains approximation, as clumps are treated as grains, although they have macroscopic sizes. They applied this approach to a clumpy infinite slab, externally illuminated on one side, and compared the results to a Monte-Carlo radiative transfer model. VD99 then refined some of the expressions of Hobson & Padman (1993) and applied them to the case of a spherical clumpy cloud, with the three types of illuminations we have discussed in Sect. 3.1.1.2: (i) central; (ii) uniform; and (iii)  external. VD99 systematically benchmarked their results with a Monte-Carlo radiative transfer code. We briefly review their results in the rest of this section.

Escape fractions for a clumpy medium. VD99 derived a series of expressions, based on Eq. (3.26), Eq. (3.29) and Eq. (3.30), but replacing grain properties by effective mega-grains properties. These analytical approximations are all summarized in Sect. 5 of VD99. We demonstrate these analytical expressions for the three types of illuminations in Fig. 3.5. Overall, they provide a good agreement with Monte-Carlo radiative transfer calculations. They are also very easy to compute. The weakest point concerns the treatment of the grain heating. The mega-grains formalism allows us to separate the absorbed fractions in the clumps and in the ICM. It thus provides different heating rates in the two phases. In Fig. 3.5, we can clearly see that the clump emission (red) is significantly colder than that of the ICM (magenta). It however does not allow us to estimate the gradient of radiation field within the ICM and within clumps. This is the most dramatic in the case of central illumination. In order to obtain a more realistic SED for this particular case, VD99 used an ad hoc prescription, assuming a power-law distribution of equilibrium grain temperatures controlled by several tuning parameters depending on the grain type.

PIC PIC
Figure 3.5: Escaping SED from clumpy spherical clouds. The three panels are the SED of a Rcl = 1 pc cloud, containing rC = 0.05 pc clumps, computed with the mega-grains approximation. The density is nICM = 1000 cm3 in the ICM and nC = 105 cm3 in the clumps (volume filling factor, fV = 20%). The grains have THEMIS optical properties and are assumed to be at thermal equilibrium. We have displayed: (i) the intrinsic stellar luminosity (T = 1.5 × 104 K; L = 3.3 × 104L ), in grey; (ii) the escaping stellar radiation, in yellow; (iii) the clump luminosity, in red; (iv) the ICM luminosity, in magenta; (v) the total escaping SED, in cyan. Licensed under CC BY-SA 4.0.
3.1.1.4 Rigorous Solutions

The radiative transfer equation (Eq. 3.16) can be solved numerically. There are two main classes of methods (Steinacker et al., 2013, for a review).

Monte-Carlo Radiative Transfer
(MCRT) consists in simulating the random walk of photons, from their sources (stars, AGNs or thermal emission from dust grains) to the outside of the region, through their multiple scatterings.
Ray-tracing
numerically solves the radiative transfer equation on a discretized grid along multiple sightlines. It is more difficult to implement than MCRT, and can be computationally more intensive. It however allows the user a better assessment of the numerical errors.

MCRT is by far the most popular method. In this section, we briefly review its principle and apply it to an example.

Setting the model. To solve the radiative transfer equation (Eq. 3.16), we need to specify the following physical ingredients.

A 3D spatial grid of dust density
has to be defined. Different coordinate systems can be chosen. In case there are steep density gradients (e.g. clumps), one has to make sure the transitions are finely enough sampled.
A 3D distribution of primary emitters,
as well as their SED needs to be defined. In this section, we will consider only stars, but AGNs can be treated the same way. It is possible to account both for: (i) discrete emission, such as an individual star or a cluster; and (ii) diffuse emission, such as unresolved stellar distribution.
The dust properties
need to include, at least: the opacity, the albedo and the asymmetry parameter. It is possible to account for a mixture of dust grains, with different compositions and sizes. The size-distribution-integrated quantities of Eqs. (2.19-2.22) are used for the transfer of stellar photons. It is however necessary to treat the thermal emission of individual species, especially if they are stochastically heated.

The principle of Monte-Carlo radiative transfer. To compute a MCRT model, we need to draw a large number of photons (typically 106 per wavelength bin), and execute the following steps. The procedure is schematically represented on Fig. 3.6.

1.
At each wavelength, photons are randomly drawn from primary emitters, proportionally to their specific intensity. The emission angle, (𝜃,ϕ), is randomly chosen, if the emitters are isotropic, which is the case for stars.
2.
The interaction probability of a photon with a dust grain is then drawn from 1 exp (τ), along the original direction of the photon. On interaction, one accounts for the probabilities of: (i) absorption (proportional to 1 ω̃); and (ii) scattering (proportional to ω̃). We could randomly choose between absorption and scattering, and thus terminate the path of the photons 1 ω̃ of the times. However, forced scattering, the way it is shown in Fig. 3.6, allows us to track the absorption and scattering probabilities at each interaction, in a numerically more efficient way. The scattering angle is randomly drawn from the scattering phase function (Eq. 1.43), setting the new direction of the photon.
3.
We iterate this process as long as needed, until the photon exits the nebula. The average number of interactions increases with the effective optical depth of the nebula. At each interaction, (1 ω̃) × Iν is absorbed by the grain, and ω̃ × Iν is scattered. After N interactions, the scattered intensity is ω̃N × I ν and the absorbed power is proportional to (1 ω̃) ×ω̃N1. For a typical ω̃ 0.5, after three scatterings, the intensity is reduced by a factor 0.13.
4.
Once all the photons at all wavelengths have been drawn, the thermal emission of all dust species within each cell can be computed. IR photons are then drawn and scattered through the nebula, the same way as stellar photons. The overall opacity and albedo are usually much lower in the IR. The computation of IR radiative transfer is thus usually much faster. In the most embedded regions, the IR radiation absorbed by dust grains can be significant. One therefore has to recompute the IR transfer a few times, until an energy balance is reached.

These steps constitute the most basic implementation of MCRT. Numerous optimizations have however been proposed in the last fifty years (e.g. Witt, 1977a,b,c; Witt & Oshel, 1977; Yusef-Zadeh et al., 1984; Whitney & Hartmann, 1992; Wood, 1997; Wood & Jones, 1997; Városi & Dwek, 1999; Baes & Dejonghe, 2001; Gordon et al., 2001; Misselt et al., 2001; Steinacker et al., 2002, 2006; Wood et al., 2008; Baes et al., 2011; Camps & Baes, 2015; Siebenmorgen et al., 2015; Natale et al., 2015; Juvela, 2019). Improvements and optimizations include: (i) massive parallelization, and the use of GPU; (ii) the production of synthetic photometric images; (iii) the treatment of polarization by scattering.

PIC
Figure 3.6: Principle of Monte-Carlo radiative transfer. The blue area represents a dust cloud. We follow the journey of a single photon emitted by a star into the cloud and scattered at different locations. A single photon will be either scattered or absorbed. However, for numerical efficiency, we consider both solutions, weighted by their probability (scattering: ω̃; absorption: 1 ω̃). The dark red arrows correspond to emitted IR photons. In principle, these photons can also be scattered and absorbed by the cloud. The arrows exiting the cloud to the right represent what would be measured by an observer. Licensed under CC BY-SA 4.0.

Numerical method to randomly draw photons. To simulate the random walk of a photon, there are two sets of random variables to draw: (i)  random interaction events; and (ii) scattering angles.

Drawing interaction events
along the path of a photon consists in drawing a path length from the following Probability Density Function (PDF):
p(l) = 1 exp 0lκρ(l)ldl. (3.34)

For simplicity, l corresponds to the path length along the direction of the photon, starting at l = 0. Drawing a variable from this distribution can be achieved the following way (e.g. Városi & Dwek, 1999). It uses the rejection method (cf. Appendix C.2.3.1). These steps are represented on Fig. 3.7.

1.
First draw a uniform random variable between 0 and 1, Θ1.
2.
Set l0 = ln (1 Θ1)κρmax, where ρmax is the maximum of the density along the path.
3.
Draw a second uniform random variable between 0 and 1, Θ2.
4.
If Θ2 ρ(l0)ρmax, then the interaction is accepted, and a new scattering angle can be drawn. On the contrary, if Θ2 > ρ(l0)ρmax, the interaction is rejected. One then needs to go back to the first step, starting from l0, this time.
Scattering angles
are drawn from the scattering phase function (Eq. 1.43; Fig. 3.8.a). If we place a spherical coordinate reference frame at the position of interaction with its z-axis aligned with the incoming direction of the photon, then the two spherical angle, (𝜃,ϕ), that will determine the new direction of the photon are drawn the following way.
1.
The new polar angle, 𝜃, is drawn by inverting the Cumulative Distribution Function (CDF) (cf. Appendix C.2.3.2). In case we use the Henyey & Greenstein (1941) phase function, the inverse of the CDF can be analytically derived (cf. Fig. 3.8.b):
F1(Θ 3) = 1 2g 1 + g2 1 g2 1 g + 2gΘ3 2 , (3.35)

where Θ3 is a uniform random variable between 0 and 1.

2.
The new azimuthal angle, ϕ, is simply drawn from a uniform distribution between 0 and 2π, because there is a symmetry of revolution around the direction of the photon.
PIC
Figure 3.7: Drawing photons in a Monte-Carlo radiative transfer model. This figure shows the dust density of a medium (in blue; normalized to its peak value, ρmax), as a function of the path length of the photon, l, across the nebula. The x-axis is multiplied by κρmax so that it corresponds to the optical depth the medium would have if the density was everywhere ρmax. Two random draws are represented. In magenta, we show a first draw, bringing the photon to l1 = ln (1 Θ1)κρmax. The interaction at l1 is rejected because Θ2 > ρ(l1)ρmax. A second similar draw brings the photon at l = l1 + l2, where the interaction is accepted. After this draw, one needs to draw the scattering angle. Licensed under CC BY-SA 4.0.
PIC PIC
Figure 3.8: Drawing scattering angles in a Monte-Carlo radiative transfer model. Panel (a) shows the Henyey & Greenstein (1941) scattering phase function (Eq. 1.43), for different values of the asymmetry parameter, g = cos 𝜃. Panel (b) shows the inverse CDF (Eq. 3.35), for different values of g. Licensed under CC BY-SA 4.0.

Example of a clumpy medium. For this manuscript, we have developed a MCRT model, following the procedure previously described. We have applied it to a spherical clumpy cloud, similar to those discussed in Sect. 3.1.1.3. The radius is Rs = 1 pc, with an ICM density, nICM = 1000 cm3. The rC = 0.05 pc clumps have a density, nC = 105 cm3, with a filling factor, fV = 20%. We assume THEMIS grain constitution, and assume all grains are at thermal equilibrium. This cloud is centrally illuminated by a T = 1.5 × 104 K star, with L = 3.3 × 104L .

The projected map
of the dust mass surface density is shown in Fig. 3.9.a. The average absorbed fraction, Pabs is shown in Fig. 3.9.b. The latter quantity is the projected average of the absorbed fraction of photons passing through the cell. This fraction is highly concentrated in the central region, as most of the power is absorbed by clumps close to the star. Those clumps are hot, whereas clumps on the outskirt are colder, because they are essentially heated by fewer photons that have been multiply scattered. Fig. 3.10 represents a few photon paths at three different wavelengths.
The total SED
of this cloud is shown in Fig. 3.11. We notice that a fraction of the dust emission is self-absorbed. This is mainly the 10 and 18 μm silicate emission of the hottest central clumps being absorbed by the outer ones. We have also compared our model to the mega-grains approximation (Sect. 3.1.1.3; in dashed blue). We see this approximation is quite good in the UV-to-NIR range. It is however discrepant in the IR. This is because there is a very strong gradient of heating conditions, seen in Fig. 3.9.b, whereas the mega-grains approximation accounts only for the total absorbed fraction in the clumps and the ICM 2.

Despite its intensive numerical requirements, MCRT is thus the most flexible and accurate way to compute the SED of an interstellar cloud. It is however important to make sure that the spatial resolution of the density grid is fine enough to resolve the shortest mean free path of photons. Otherwise, we would be smearing out a potential sub-grid temperature gradient. For the present model, we have 0.01 pc cells for lmean(U) 0.005 pc.

 To accurately compute a radiative transfer model, it is necessary to resolve scale-lengths of the order of the mean free path of UV photons.

PIC PIC
Figure 3.9: Spatial distributions of the clumpy radiative transfer model. Panel (a) shows the projected dust mass surface density of the cloud. Panel (b) shows the average projected absorbed fraction, Pabs. The star is located at the center of the cloud. Notice that most of the power is absorbed in the central region. Licensed under CC BY-SA 4.0.
PIC PIC PIC
Figure 3.10: Random photon path within a clumpy medium. The three grey-scale images represent the projected column density of Fig. 3.9.a. We have overlaid the random path of 10 photons before they exit the cloud, at three different wavelengths. These draws are those computed by the MCRT model. For a given wavelength, the different colors represent different photons. These figures demonstrate the fact that, at short wavelength, the mean free-path is shorter, resulting in a larger number of scatterings. Licensed under CC BY-SA 4.0.

PIC

Figure 3.11: Total SED from the clumpy radiative transfer model. The yellow curve represents the intrinsic SED of the central star. The red curve is the intrinsic emission (non self-extincted) of the dust emission of the whole cloud. The grey curve shows the total escaping SED of the cloud. We have overlaid the mega-grains approximation (Sect. 3.1.1.3), in dashed-blue, for comparison. Licensed under CC BY-SA 4.0.

3.1.2 Approximate Treatments of the Mixing of Physical Conditions

Ideally, each time we study a Galactic region or a galaxy, we should solve the radiative transfer equation (Eq. 3.16). This is however, most of the time, impossible, because of the lack of constraints on the actual 3D structure of the region. Even if we have a collection of high-angular resolution multiwavelength images of our object, the matter and stellar distributions along the sightline have to be inferred. This inference is possible when the large-scale geometry of the object is quite regular, for instance: a protostellar disk and its jet, or a galactic disk and its bulge, etc. We will discuss the MCRT modeling of disk galaxies in Sect. 3.1.3.2. Otherwise, most often, we need to adopt empirical approaches that allows us to constrain the dust properties, despite our uncertainty of the spatial structure of the region.

3.1.2.1 The Historical Model: the MBB

The MBB (cf. Sect. 1.2.4.1) is historically the most widely-used dust SED model. It is controlled by the three following parameters (Eq. 1.70 and Eq. 1.72).

1.
The dust mass, Mdust, is a scaling parameter.
2.
The equilibrium temperature, T, controls the emission peak wavelength.
3.
The emissivity index, β (Eq. 1.72), controls the long-wavelength slope of the SED. This is demonstrated in Fig. 3.12.a.

Its physical assumptions are simplistic: (i) the IR emission is optically thin; (ii) the dust is made of a single species of grains at thermal equilibrium with the ISRF; and (iii) the opacity is a power-law. By inferring both T and β, this model is designed to constrain both the dust excitation and its optical properties. This model was popularized by Hildebrand (1983), in the IRAS days (cf. Sect. 2.1.2). At the time, it was well adapted, being a simple, but still physical model, with only three parameters to fit four broadbands (the four IRAS bands at 12, 25, 60 and 100 μm). It has however several limitations that are often disregarded in the literature.

The mixing of physical conditions
is not accounted for by the MBB. It means that the SED fit of a complex region, containing a gradient of temperatures, is biased (e.g. Juvela & Ysard, 2012; Galliano et al., 2018). This is demonstrated in Fig. 3.12.b. The grey-filled curve is the SED coming from a power-law distribution of temperatures (T = 15 to 60 K; the color curves). The intrinsic emissivity index of the grains making up this region is β = 1.79. Yet, fitting such a SED with a single MBB leads to a compromise temperature (T = 45 ± 8 K) and a systematically lower emissivity index, β = 1.12 ± 0.30. This is because a gradient of temperature tends to flatten the FIR slope, similarly to a lower β value.

 The emissivity index derived from a single MBB fit is always lower than its true, intrinsic value.

Stochastic heating
is formally equivalent to a gradient of temperatures (cf. Sect. 1.2.4.3). Stochastically heated grains dominate the emission at MIR wavelengths (cf. Fig. 2.27). A single MBB is thus biased at short wavelengths by small grains, the same way it is biased at long wavelengths by cold dust. Solving this issue by fitting a linear combination of two or three MBBs can palliate this problem, but is usually not sufficient. In addition, fitting an out-of-equilibrium emission with equilibrium grains is physically incorrect. It renders the interpretation of the parameters of the hot MBB unreliable.
Realistic opacities are more complex
than a power-law. Laboratory data show that emissivity indices of actual materials are wavelength-dependent quantities, β(λ) (cf. e.g. Fig. 2.22). Additionally, the somehow arbitrary choice of the two other parameters in Eq. (1.72) – the reference wavelength, λ0, and the scaling of the opacity at this wavelength, κ0 – have dramatic consequences on the derived dust mass. It is thus important to limit the potential range of variations of these parameters. The Kramers-Kronig relations (Sect. 1.2.1.6) impose that β 1, and most compounds studied in the laboratory have β < 2.5 (cf. e.g. the Jena database). We recommend calibrating κ0 on laboratory data or on well-constrained dust models (such as Eq. 2.26). Contrary to what Hildebrand (1983) recommended, it is probably better to choose a reference wavelength, λ0, around the peak of the FIR SED (between 100 and 300 μm), rather than in the submm. This way, variations of β will impact only moderately the derived dust mass.

 A MBB fit infers parameters whose physical meaning is difficult to assess.

The noise-induced degeneracy
between T and β is well-documented (e.g. Shetty et al., 2009; Kelly et al., 2012; Galliano, 2018). A false negative correlation arises when a series of MBB fits are performed with standard least-squares, maximum likelihood, or non-hierarchical Bayesian methods. This degeneracy prevents to explore the potential correlation of these parameters that laboratory data and solid-state models suggest (e.g. Mennella et al., 1998; Meny et al., 2007). Luckily, hierarchical Bayesian methods are efficient at removing false, noise-induced correlations (e.g. Kelly et al., 2012; Galliano, 2018). Chap. 5 is entirely devoted to fitting techniques.
PIC
Figure 3.12: MBB fitting. Panel (a) demonstrate the effect of the emissivity index, β (Eq. 1.72), on the SED of a MBB. The temperature is T = 30 K, with κ(λ0 = 250μm) = 0.64 m2/kg (Eq. 2.26). Panel (b) illustrates the limitation of the isothermal assumption. The black error bars are synthetic observations (noise has been added). They are sampling the grey-filled curve, which is the integral of MBBs (color curves) times a power-law distribution of temperatures (Tmin = 15 K, Tmax = 60 K, and index 4). These MBBs have THEMIS optical properties (β = 1.79; Eq. 2.26). Yet the fitted value (black dashed curve) is significantly lower: β = 1.12 ± 0.30. The fitted temperature falls in the middle of the range, toward the high end: T = 45 ± 8 K. Licensed under CC BY-SA 4.0.

3.1.2.2 A Phenomenological, Composite Approach

Dust models provide useful frameworks to model SEDs (Sect. 2.3). Without the possibility to compute the radiative transfer, we are however facing the problem of the mixing of physical conditions. A prescription, proposed by Dale et al. (2001), has proven to be a powerful solution to this issue. It consists in assuming that the dust mass is distributed in regions with different starlight intensities, U, following a power-law:

1 Mdust dMdust dU = Uα× 1 α (U + ΔU)1α U1α  if α1 1 ln (U + ΔU) ln U  if α = 1  for U U U+ΔU. (3.36)

The idea is that the shape of the observed SED is used to constrain this distribution of ISRFs, assuming a dust mixture constitution. By lack of a better term, we call this approach the composite model. The free parameters are:

It thus provides a way to account for the potential complexity of the region without having to model the radiative transfer. The model SED is then simply:

Lνdust(λ) =UU+ΔU𝜖 ν(λ,U)dMdust dU dU, (3.37)

where 𝜖ν(λ,U) is the monochromatic emissivity of the dust model, exposed to a single starlight intensity, U (Fig. 2.24). Fig. 3.13.a shows an example of a SED fit, using Eq. (3.36). We have a added a free-scaling black body, at T = 30, 000 K, to account for the stellar continuum that may be contaminating the MIR photometric bands. The composite approach is flexible enough to be usable in a diversity of environments. Dale et al. (2001) lists several simple geometries for which Eq. (3.36) is actually the solution. It is also adapted to more complex ISM topologies. For instance, Fig. 3.13.b shows the dust mass distribution as a function of U, for each cell in the MCRT simulation of Sect. 3.1.1.4. Despite the complex, clumpy structure of this cloud, it can be reasonably well-approximated by a power-law (shown in red).

PIC PIC
Figure 3.13: Phenomenological mixing of physical conditions. Panel (a) shows an example of SED fit using the composite approach (Eq. 3.36). The black error bars represent synthetic observations (noise has been added). These observations have been fitted with the total model represented by the dark grey line. This total model is the sum of a stellar (in yellow) and a dust (in light grey) components. The dust component is the integral of dust models illuminated by different U (Eq. 3.36). The rainbow curves represent several bins of ISRF. The sum of these rainbow curves is the light-grey-filled curve. The blue curve in panel (b) shows the distribution of dust mass per bin of ISRF intensity, U (Sect. 1.2.4.2), for each cell in the MCRT simulation presented in Sect. 3.1.1.4. The value of U has been derived from the grain equilibrium temperature, using Eq. (1.83). The red line shows a power-law approximation of this distribution. With the parametrization of Eq. (3.36), it would correspond to: U = 0.03, ΔU = 104 and α = 0.5 + 1 = 1.5 (the +1 comes from taking the derivative of Mdust(U) to get Eq. 3.36). Licensed under CC BY-SA 4.0.

The average starlight intensity. The parameters of Eq. (3.36) do not have a very clear physical meaning. Besides, they are often degenerate: the uncertainties on the three parameters U, ΔU and α are strongly correlated. It can be more efficient to quote the average of the distribution:

U 1 MdustUU+ΔUU dMdust dU dU = 1 α 2 α × U + ΔU2α U 2α U + ΔU1α U1α  if α1&α2 ΔU ln U + ΔU ln U  if α = 1 ln U + ΔU ln U U1 U + ΔU1  if α = 2. (3.38)

This parameter quantifies the average starlight intensity, heating the bulk of the dust mass. It is the equivalent of the equilibrium temperature of a MBB, as it controls the peak emission wavelength, except that it accounts for the mixing of physical conditions and the stochastic heating of small grains. The Total InfraRed (TIR) luminosity, LTIR, can be expressed:

LTIR 3μm1000μmL λ(λ)dλ 0L νdust(ν)dν = 𝜖 × M dust ×U, (3.39)

where the constant 𝜖 𝜖νdν is the bolometric emissivity of the dust model. For the THEMIS mixture, heated by the Mathis et al. (1983) ISRF, it is 𝜖 = 221LMU (Table 2.7).

Constraining the dust properties. The composite approach of Eq. (3.36) allows us to constrain parameters that are not extremely sensitive to radiative transfer effects (i.e. to variations of the local intensity and spectral shape of the ISRF; cf. e.g. Galliano et al., 2018, 2021, for a discussion).

The dust mass,
Mdust, is dominated by large grains (cf. Table 2.3). The heating of these large grains is sensitive only to the power they absorb, and not to the spectral shape of the ISRF, contrary to small grains (cf. Sect. 1.2.4.3). The assumption of a constant ISRF shape therefore does not significantly bias Mdust.
The average starlight intensity,
U (Eq. 3.38), is empirically constrained by the shape of the FIR emission, dominated by large grains. This quantity is thus not particularly biased, similarly to Mdust. In addition, Eq. (3.39) tells us that LTIR MdustU. Yet, LTIR is a weakly model-dependent quantity, as it is simply the integral of the SED model passing through the observed fluxes. LTIR and Mdust being reliable, U has to be.
The mass fraction of PAH,
qPAH (or the mass fraction of small a-C(:H), qAF, for the THEMIS model), discussed in Sect. 2.3.3.1, controls the strength of the MIR aromatic features (cf. Sect. 2.2.2.1). These features, being carried by small, stochastically heated grains, are sensitive to the spectral shape of the ISRF (cf. Fig. 2.27.a). This is demonstrated in Fig. 3.14. We have considered two extreme ISRFs (Fig. 3.14.a). The bias on the aromatic feature emission is at most a factor 1.8 (Fig. 3.14.b).

 Eq. (3.36) provides acceptable estimates of Mdust, U and qPAH (or qAF).

PIC PIC
Figure 3.14: Effect of the ISRF hardness on the SED. Panel (a) represents different ISRFs: (i) the Solar neighborhood (red;  Mathis et al., 1983); and (ii) a B star (T = 30000 K; blue;  Lanz & Hubeny, 2007). We have represented the spectral range of ionizing photons in yellow. Both ISRFs are normalized so that U = 1 (Eq. 1.82). Panel (b) shows the SEDs of the THEMIS model, heated by the two ISRFs. We show the SED for different values of U, in order to obtain the same FIR SED, as this parameter is inferred by the fitter. The SEDs are normalized by the FIR luminosity, LFIR, between λ = 60 and 200 μm. The B star heating leads to an increase in the λ = 6 9μm emission by a factor 1.8, compared to the Solar neighborhood. Licensed under CC BY-SA 4.0.

Limitations of the composite approach. Using Eq. (3.36) to model typical broadband SEDs of Galactic regions and nearby galaxies is an efficient method. However, as any approximation, it has some limitations.

Variation of the grain constitution
within the observed region is expected, as dust evolves with ISRF intensity and ISM density (cf. Chap. 4, which is devoted to dust evolution). The composite approach accounts for the fact that there are variations of the physical conditions within the region. However, assuming that the dust constitution is homogeneous biases the derived properties. The most problematic aspect is certainly the variation of the overall FIR opacity. If there is significant mantle accretion in dense regions, we might be mixing together regions with different κ (cf. Fig. 1.21). This is unavoidable and this has to be pondered when discussing the modeling results.
The degeneracy
between the grain size and ISRF distributions prevents constraining the former. This is because a MIR excess due to an enhancement of the small grain emission looks similar to the presence of hot equilibrium grains (such as compact H II regions) Fig. 3.15 demonstrates this degeneracy. It shows the fit of the same synthetic observations, either varying the fraction of small grains, or the mixing of ISRFs. Additional constraints, on the geometry of the region and its radiation field, are necessary to attempt breaking this degeneracy (cf. Sect. 3.1.3.4).
PIC
Figure 3.15: Degeneracy of the grain size and ISRF distributions. The two panels show the same synthetic observations (black error bars) fitted by two different models (Galliano et al., 2018). Panel (a) represents a fit by playing on the proportions of grains of different sizes: PAHs, VSGs and BGs. Panel (b) represents a fit using Eq. (3.36). In both cases, we have used the THEMIS dust mixture. Licensed under CC BY-SA 4.0.

An alternative distribution. The starlight distribution of Eq. (3.36) is not the only possible one. Draine & Li (2007) proposed the following:

1 Mdust dMdust dU DL07 = γ × 1 Mdust dMdust dU Eq.(3.36) + (1 γ) × δ(U U), (3.40)

which simply is the linear combination of the power-law distribution of Eq. (3.36), with a Dirac distribution centered at U = U, fixing α = 2 and U + ΔU = 106. The power-law distribution is supposed to account for star-forming regions, with a large gradient of starlight intensity, and the Dirac represents the diffuse ISM, uniformly illuminated. The parameter γ controls the relative weight of these two components. This distribution was elaborated when Spitzer data were being analyzed. There was no coverage beyond λ = 160μm. There was thus no constraint on the Rayleigh-Jeans regime of the SED. The role of the Dirac component was to fit the SED peak, avoiding the dust mass to explode by lack of constraint on the cold dust distribution. This parametrization however became problematic when Herschel data arrived. The FIR-submm slope of the observed SED can indeed not be fitted as well with this model (Eq. 3.40) as with the composite approach (Eq. 3.36). This is because the model’s slope is the intrinsic slope of the grain mixture. This is demonstrated in Fig. 3.16.a. In Fig. 3.16.b, we see that the starlight intensity distribution fit can not go down as low as the composite model. The Dirac component fits the FIR peak with a compromise U. This is reminiscent of the discussion we had about MBB fits, in Sect. 3.1.2.1. This was demonstrated by G21, who compared different approaches by fitting the SEDs of about 800 galaxies of the DustPedia project (Davies et al., 2017) and Dwarf Galaxy Sample (DGS;  Madden et al., 2013). Fig. 3.17 compares the composite approach, as a reference, to the following models (see the complete discussion in  Galliano et al., 2021).

A MBB with β free
infers dust masses that are, on average over the whole sample, consistent with the composite approach (Fig. 3.17.a). However, we see there is a bias as a function of the mass of the galaxy. The MBB approach tends to overestimate the mass of late-type galaxies (mostly the high-mass objects), and underestimates the mass of early-type galaxies. This is because the SED of early-type galaxies is fitted with a smaller β (thus a lower temperature, and a lower dust mass).
A MBB with β fixed
tends to underestimate the dust mass on average. In Fig. 3.17.b, the median of the ratio is 0.85. This is the effect we have discussed in Sect. 3.1.2.1: the mixing of physical conditions is fitted by a compromise temperature, underestimating the coldest dust.
The Draine & Li (2007) distribution
(Eq. 3.40) results in a similar bias as the MBB. For the distribution in Fig. 3.17.c, the median of the ratio is 0.71, even lower than in the case of the MBB. The reason is the same: the uniformly illuminated component dominating the FIR-submm SED can not account for the distribution of cold dust.
PIC PIC
Figure 3.16: SED fit with the Draine & Li (2007) starlight intensity distribution. The black error bars in panel (a) are synthetic observations. They are fitted using Eq. (3.40): (i) the power-law component in magenta; (ii) the Dirac component in blue; and (iii) an additional stellar continuum in yellow. The corresponding starlight intensity distribution is shown in panel (b), with the same color code. We have overlaid, in green, a composite model fit to the same synthetic observations. Licensed under CC BY-SA 4.0.

PIC

Figure 3.17: Comparison of different SED models. The data are from the SED modeling of the DustPedia and DGS nearby galaxies by Galliano et al. (2021, 800 galaxies of all types). The x-axis of the left panels shows the dust mass derived by fitting the SED of each galaxy using the composite approach (Eq. 3.36). The y-axis shows the ratio of the x-axis and the dust mass estimate using one of the three alternate models: (i) MBB; (ii) MBB with fixed β = 1.79; (iii) Draine & Li (2007) distribution. Each point with an error bar corresponds to one particular galaxy. The yellow line shows the 1:1 ratio, and the dashed red line, the median of the sample. The right panels show the histogram of the distribution. Licensed under CC BY-SA 4.0.
3.1.2.3 Panchromatic Empirical SED Models

Several codes in the literature model the panchromatic 3 SED of galaxies, with a simplified treatment of the radiative transfer (e.g. Silva et al., 1998; Charlot & Fall, 2000; Galliano et al., 2008a; da Cunha et al., 2008; Boquien et al., 2019; Fioc & Rocca-Volmerange, 2019). They include the emission by stellar populations, in addition to dust SEDs.

Stellar SEDs. Stars of different masses have different spectra and lifetimes. This is shown in Fig. 3.18, in the form of a Hertzsprung-Russell diagram. Massive stars have: (i) the highest effective temperatures, their SEDs peaking in the UV; (ii) the highest luminosities; and (iii) the shortest lifetimes, of only a few million years. This is the opposite for low-mass stars: their SED peaks in the NIR, and they live longer than several hundred million years. These characteristics have a profound impact on the variation of stellar SEDs with time. The intrinsic emission of a stellar population can be simulated using evolutionary synthesis (e.g. Fioc & Rocca-Volmerange, 1997). This approach follows, at each time step, the formation and evolution of stars with different masses, m. When these stars are born, all populations contribute to the SED. It is shown as the magenta curve in Fig. 3.19. This initial SED is dominated by massive stars and peaks in the UV. Then, as the most massive stars, which also have the shortest lifetime, die, their contribution to the SED is suppressed. Consequently, when these stars disappear, the UV-side of the SED decreases. After several hundred million years, the SED peaks in the NIR, as it originates only in low-mass stars (orange and red curves in Fig. 3.19). There is also a drastic evolution of the emissivity as a function of time, as low-mass stars are significantly less luminous (cf. Fig. 3.18). To compute such synthetic spectra, the following quantities need to be defined.

We will more extensively discuss these quantities in Sect. 4.3, when modeling the chemical evolution of galaxies. We have summarized in Table 3.2 the main properties of the different stellar classes.

PIC
Figure 3.18: Stellar isochrones. This figure shows theoretical tracks of individual star evolution from the model of Schaller et al. (1992), at initial metallicity, Z = 0.008. This representation is known as a Hertzsprung-Russell diagram. Stars of a given mass, m (a given color), start from the Zero Age Main Sequence (ZAMS; grey) and evolve along their Main Sequence (MS) track until their death, after a time τ(m). We have highlighted the late evolutionary stages: (i) red supergiants for massive stars; and (ii) AGB for Low- and Intermedate-Mass Stars (LIMS). Licensed under CC BY-SA 4.0.

PIC

Figure 3.19: Evolution of stellar SEDs as a function of time. The different curves show the evolution, as a function of time, of the SED of a stellar population produced by an instantaneous burst, at Solar metallicity. We have assumed the Salpeter (1955) IMF. These models were generated using the code of Fioc & Rocca-Volmerange (1997). Notice, in particular, the fast decrease of ionizing photons with time. Licensed under CC BY-SA 4.0.






Effective temperature, Teff Initial mass, m Initial luminosity, L Lifetime, τ(m)





O

30000 K 16M 30000L 12 Myr

B

10000 30000 K 2.1 16M 25 30000L 12 Myr1 Gyr

A

7500 10000 K 1.4 2.1M 5 25L 1 3 Gyr

F

6000 7500 K 1.04 1.4M 1.5 5L 3 10 Gyr

G

5200 6000 K 0.8 1.04M 0.6 1.5L 10 25 Gyr

K

3700 5200 K 0.45 0.8M 0.08 0.6L

M

2400 3700 K 0.08 0.45M 0.08L





Table 3.2: Basic properties of the main stellar classes.

Putting everything together. Empirical panchromatic SED models usually include the following physical ingredients.

One or several stellar populations
are modeled. Their escaping radiation is fitted to the UV-to-NIR SED, allowing the user to constrain: the age, the SFH timescale, the SFR and the total stellar mass. The attenuation 4 is accounted for, by assuming simple topologies, such as those discussed in Calzetti et al. (1994). The intrinsic stellar power, L, is derived from the energy balance: L = LUV-NIR + LTIR, where LUV-NIR is the escaping UV-to-NIR power. This equation simply states that the intrinsic power emitted by stars is either escaping the galaxy in the UV-to-NIR range, or has been absorbed by dust. It implicitly assumes that the emission from the galaxy is isotropic, which is not necessarily the case for disk galaxies.
Dust emission
is modeled, either by solving the radiative transfer equation (Eq. 3.16) in a simple geometry, or by adopting a distribution of starlight intensities (cf. Sect. 3.1.2.2). The dust properties are not always consistent with those used to account for attenuation.
Nebular emission,
in the form of lines or continuum, can be included. In case of star-forming galaxies, it is dominated by H II regions (e.g. Charlot & Longhetti, 2001). Some models also include the gaseous emission from AGNs. These tracers are used to refine some of the diagnostics.

We illustrate this approach with the model of Galliano et al. (2008a), applied to two galaxies, in Fig. 3.20. This particular model separates the emission from H II regions, which are powered by massive ionizing stars (cyan). The magenta curve shows the escaping radiation from H II regions. It includes the dust emission and the free-free continuum. The escaping radiation from H II regions, as well as non-ionizing field stars (yellow), heat the neutral ISM (red). The degeneracy between both dust components was solved by using radio observations to constrain the level of the free-free emission. The gas density in H II regions, impacting the equilibrium temperature of large grains, is constrained by the MIR continuum. The remaining emission is thus assumed to come from the neutral ISM. We will discuss in more detail the results of this model, in Sect. 4.3.

PIC
Figure 3.20: Multiphase SEDs of galaxies. This figure shows the SED of two galaxies, modeled by Galliano et al. (2008a): the prototypical starburst, M 82, and the lowest metallicity galaxy, I Zw 18. The black error bars represent the observed fluxes. The total SED model is shown in grey. The stars are split in two populations: (i) young, ionizing stars that are in H II regions; and (ii) older, non-ionizing stars that are in the rest of the galaxy. We show the intrinsic emission of both stellar components. The dust emission originates in: (i)  H II regions, heated by ionizing stars; and (ii) the neutral ISM, heated by non-ionizing stars and by the escaping radiation from H II regions. Licensed under CC BY-SA 4.0.

Limitations of empirical panchromatic models. The approach we have just described has several advantages. In a single fit, it provides estimates of the SFR, the age of the stellar populations, the stellar mass, and the dust properties. Its major limitation however resides in the sensitivity of its results to the assumed ISM topology. The ISM indeed has a fractal structure, with several orders of magnitude of contrast density (e.g. Combes, 2000). The optical depth from the model thus probably underestimates the total dust column density (cf. Sect. 3.1.1.3). In addition, the extinction and emission properties of these models are usually not consistent. The modeling of the microscopic grain constitution and of their macroscopic spatial distribution can differ from one side of the electromagnetic spectrum to the other. For instance, assuming a Calzetti et al. (1994) attenuation law and taking a mean radiation field to account for dust heating is virtually equivalent to decoupling the extinction and emission. Finally, the consistency brought by modeling all multi-wavelength tracers at once, which is a priori positive, leads to propagating the systematic uncertainties, due to arbitrary choices of ISM topologies, to parameters that could have been considered independent of these effects, if they had been modeled separately (Mdust, U, qPAH; Sect. 3.1.2.2).

3.1.2.4 The Matryoshka Effect

The main limitation of dust studies lies in the near impossibility to constrain at the same time the microscopic dust properties and their sub-pixel macroscopic spatial distribution. All the approaches we have discussed in this chapter face this problem. It can be illustrated with what Galliano et al. (2018) called the matryoshka effect. This empirical effect originates in the impact of the spatial resolution of the observations on the constrained parameters.

Demonstration on the LMC. Fig. 3.21 demonstrates the effect with the modeling of the dust mass in a strip covering one fourth of the LMC, by Galliano et al. (2011). The different images on top show the spatial resolution that is used. The first image is one single large pixel. The second one is divided in four pixels, and so on. The curve in the bottom panel of Fig. 3.21 shows the dust mass in the strip derived by summing all the pixels, at each resolution. We see that this mass increases with spatial resolution, until reaching a plateau around 10 pc. For this particular region, the discrepancy with the global mass is about 50%. Galliano et al. (2011) interpreted this effect by noticing that the cold dust, which accounts for most of the mass and, at the same time, is the least luminous, is diluted into the warm dust emission when we sum all the regions together. With spatial resolution however, we can better separate the bright and cold regions. It is thus reasonable to assume that the most accurate estimate of the dust mass is the one obtained at the highest resolution. This assumption is confirmed by the fact that the length-scale at the growth curve plateau ( 10 pc) corresponds roughly to the mean free-path of a U-band photon at a density nH 100 cm3 (Table 3.1; the LMC has a half-Solar metallicity; Pagel, 2003). This is the typical density of the CNM and diffuse H2 phase (cf. Table 3.6). It is possible that, if we could increase the resolution, we would see another plateau around nH 104 cm3 ( 0.1 pc), corresponding to dense H2 clouds.

PIC PIC PIC PIC PIC PIC PIC PIC PIC PIC
PIC
Figure 3.21: Matryoshka effect demonstrated on the LMC. On top are nine SPIRE500μm images of a strip covering about one fourth of the LMC, observed by Meixner et al. (2010). These images are rebinned at different spatial resolutions (indicated on the top). The bottom panel shows the dust mass derived by modeling the SED of each pixel of these images and summing them to derive the total mass in the strip (Galliano et al., 2011). It is normalized by the mass obtained when modeling the global (i.e. one pixel) SED of the strip, Mdustglobal. Licensed under CC BY-SA 4.0.

Generalization. This effect has been independently confirmed by Galametz et al. (2012), Roman-Duval et al. (2014) and Aniano et al. (2020), although it is less important if the maximum spatial resolution is not as high as ours.

 The dust mass derived at high spatial resolution is always larger than its global estimate.

3.1.3 Application to Nearby Galaxies

We now review the application of SED models to observations of nearby galaxies, aimed at constraining the grain properties in different environments. We illustrate the different aspects using our own projects and collaborations.

3.1.3.1 The Different Types of Galaxies

There is a wide diversity of galaxy types. Several systems of classification have been developed, through the years. In particular, the Hubble-de Vaucouleurs system, although outdated, is still used nowadays.

The outdated galaxy morphological classification. The most famous observational system of morphological classification is the Hubble tuning fork or Hubble-de Vaucouleurs diagram, represented in Fig. 3.22. It was originally developed by Hubble (1936), and refined by de Vaucouleurs (1959). It is based on the morphological characteristics of galaxies in the visible range: presence and thickness of spiral arms, bars, rings, etc. There are essentially three classes of galaxies (left, center and right parts of Fig. 3.22): (i) ellipticals, also called Early-Type Galaxies (ETG); (ii) spirals, also called Late-Type Galaxies (LTG); and (iii) irregulars and dwarf spheroidals. There are sub-categories with abstruse notations (SAa, E2, etc.) that would be a waste of time to describe here. Overall, this is a mid-XXth-century empirical classification, based on stellar dynamics, that does not take into account the IR information (especially the SF activity), nor the radio properties (presence of an AGN). Recently, with the DustPedia collaboration (Davies et al., 2017), we explored the sensitivity of dust properties to galaxy types (e.g. Davies et al., 2019; Bianchi et al., 2018; Nersesian et al., 2019). We did not find any clear systematic differences between adjacent sub-categories in Fig. 3.22, and we found a large scatter within each class. There are overall trends between the three main categories, because they are linked to the gas fraction, metallicity and stellar populations. We will discuss those throughout this manuscript. The terminology “late/early-type” was introduced to denote the evolution of galaxies along the sequence. We now know that the sequence is reversed: early-type galaxies are the oldest objects, and late-types, the youngest ones. In addition, the most numerous galaxies in the local Universe, dwarf galaxies are under-represented in this diagram. They are part of the irregulars, which is a category by default. Finally, this Hubble-de Vaucouleurs classification was based on the local Universe, while galaxies at high redshift can exhibit different morphologies, such as clumpy chains and tadpoles (e.g. Elmegreen, 2015). We have developed an alternate, non-parametric classification, taking into account IR morphology (Baes et al., 2020). It emphasizes the clumpy nature of the dust distribution in local galaxies.

 Stellar morphology is not particularly relevant to ISM studies.

PIC
Figure 3.22: Hubble-de-Vaucouleurs galaxy morphology diagram. Credit: A. Ciccolella / M. De Leo, Wikipedia, licensed under CC BY-SA 3.0.

Galactic properties that matter to ISM studies. A few global quantities, such as the metallicity, the SFR or the gas fraction are more relevant to assess the general properties of the ISM of a galaxy. The three main categories of the Hubble-de Vaucouleurs diagram can be characterized the following way.

Irregular galaxies
(e.g. Fig. 3.23.a) can contain large amounts of atomic gas that typically extend to twice their Holmberg radius 5 (e.g. Huchtmeier et al., 1981).
At Solar metallicity,
irregulars are rich in dust. Their visible and MIR scale-lengths are very similar (Hunter et al., 2006). The dust emission in irregular galaxies is clumpy and warm, with hot dust and UIB emissions mostly observed towards bright H II regions. This suggests that massive stars are a major source of heating in these environments (e.g. Hunter et al., 2006).
At low metallicity,
irregulars are dwarf galaxies. Dustiness increases with metallicity (cf. Sect. 4.3; Rémy-Ruyer et al., 2014; Galliano et al., 2021). The ISM in these objects is less dusty and thus, more transparent. Similarly to irregular galaxies, massive stars are a major source of heating in these objects (e.g. Walter et al., 2007), and they are permeated by SN-triggered shock waves (e.g. Oey, 1996; Izotov et al., 2007). Finally, these galaxies exhibit large H I envelopes. The IR emitting region can correspond to only 20 30% of the total mass of the system (e.g. Walter et al., 2007). A particularly important type of dwarf galaxies are Blue Compact Dwarf Galaxies (BCD). These galaxies are actively star forming. As their name indicates, they have blue optical colors, because they contain several SSCs, rich in young massive stars, and they are weakly attenuated by dust.
Late-type galaxies
(e.g. Fig. 3.23.b) have a roughly Solar metallicity and an ISM accounting for 10 30% of their baryonic mass.
Scale-lengths
of disk galaxies are of the order of a few kpc (up to 10 kpc). Their diameter tends to be systematically smaller in the MIR than in the visible (Malhotra et al., 1996), whereas the arm/inter-arm contrast is larger in the MIR than in the visible (Vogler et al., 2005). This is also seen in 12CO(J =1 0) 2.6mm, Hα656.3nm or radio continuum (Sauvage et al., 1996; Walsh et al., 2002; Vogler et al., 2005). This is because these different tracers are indicators of star formation activity, which is enhanced along the spiral arms. In the IR, the disk scale-length increases with wavelength (Hippelein et al., 2003), and is larger in the FIR than in the visible (Tuffs et al., 1996; Alton et al., 1998; Haas et al., 1998; Davies et al., 1999; Trewhella et al., 2000; Fritz et al., 2012; Casasola et al., 2017). This FIR colour gradient observed in the disk suggests that part of the FIR emission arises from grains heated by the radially decreasing diffuse ISRF, the extended parts of the disk being cold (Block et al., 1994; Stevens et al., 2005; Hinz et al., 2012). FIR scale-lengths do not vary strongly as a function of galaxy type and are on average 10% larger than the stellar scale-lengths (e.g. Muñoz-Mateos et al., 2009; Hunt et al., 2015).
Scale heights
of disk galaxies are typically of the order of a few hundred parsecs. Outside our Milky Way, edge-on galaxies are ideal objects to constrain this parameter (e.g. Xilouris et al., 1999). Radiative transfer codes are robust tools to derive such structural parameters (cf. Sect. 3.1.3.2).
Early-type galaxies
(e.g. Fig. 3.23.c) possess very little dust: the average dust-to-stellar mass ratio is 50 times lower than that of spiral galaxies (Smith et al., 2012; Galliano et al., 2021). Dust lanes are, however, commonly detected in elliptical galaxies (e.g. Sadler & Gerhard, 1985; Gomez et al., 2010). Jura et al. (1987) for instance found that half of nearby ellipticals are detected at IRAS wavelengths. Smith et al. (2012) found that elliptical galaxies detected at 250 μm tend to have higher X-ray luminosities. This was further explored by G21. We will come back to this point in Sect. 4.2.2.2.
(a) Irregular, dwarf (b) Spiral (c) Elliptical
(I Zw 18) (M 33) (CentaurusA)
PIC PIC PIC
Low metallicity ( 135Z) Solar metallicity ( 12Z) High metallicity ( 3Z)
Figure 3.23: Visible-range image of three nearby galaxies. Panel (a) shows an image of the extremely low-metallicity, dwarf galaxy, I Zw 18, obtained with the instrument ACS onboard the HST (Aloisi et al., 2007). Panel (b) shows the spiral galaxy, M 33. Star-forming regions, traced by the Hα656.3nm line, are shown in red. Panel (c) shows the elliptical galaxy, Centaurus A. It has a warped dust lane and an AGN. Credit: (a) Cignoni & Tosi (2009), licensed under CC BY-SA 3.0; (b) Image Data: Subaru Telescope (NAOJ), Hubble Space Telescope – Image Processing: Robert Gendler – Additional Data: BYU, Robert Gendler, Johannes Schedler, Adam Block – Copyright: Robert Gendler, Subaru Telescope, NAOJ, with permission from Robert GENDLER; (c) courtesy of ESO, licensed under CC BY 4.0.

3.1.3.2 Large-Scale Dust Distribution in Galaxies

The dust spatial distribution, that is the value of various dust parameters in different regions or pixels of a galaxy, can be determined in the MW and nearby galaxies. This determination however requires good quality, homogenized multi-wavelength images of the studied objects. This is therefore significantly more complex than modeling the SED of a point source.

Homogeneous multi-wavelength data sets. To accurately model SEDs of galaxies, the observed fluxes must originate, at each wavelength, in the same region, and must trace only the emission we are modeling (i.e. the dust emission, the escaping stellar emission, etc.). The following artifacts can be encountered.

Contaminations
in the telescope beam can have several origins. Most of them are displayed in Fig. 3.24.
Foreground zodiacal emission
originates in dust grains from the Solar system (e.g. Fixsen & Dwek, 2002; Reach et al., 2003; Rowan-Robinson & May, 2013; Krüger et al., 2019). Its intensity depends greatly on the angle above the planetary disk. It is prominent in the MIR (cf. Fig. 3.24). The zodiacal emission is quite homogeneous at the typical angular scale of nearby galaxies (a few degrees or less). It can thus be efficiently subtracted, using the emission outside the target. Otherwise, several models compute its synthetic spectrum (e.g. Krüger et al., 2019).
Foreground Galactic emission
originates in dust grains from cirrus clouds within the MW. Its spectrum does not vary significantly with pointing direction, as the Galactic ISRF is relatively homogeneous. However, its total intensity scales with the column density of the ISM between the observer and the target galaxy. The H column density is the lowest at high Galactic latitude (NH 1024 H/m2). The fractal structure of interstellar clouds (e.g. Elmegreen & Falgarone, 1996) results in a high degree of structure at small angular scales. In other words, it is difficult to subtract this emission using off-target areas. This contamination can not always be subtracted. Surface brightness being independent of distance, the emission from the diffuse ISM of nearby galaxies is similar to that of the MW, making the study of the former particularly challenging.
Background galaxy emission
constitutes the CIB (e.g. Dole et al., 2006). Its SED looks similar to the diffuse Galactic ISM (cf. Fig. 3.24). Spatially, it is granular, as it is the sum of numerous point sources. It is difficult to accurately subtract. It is thus another component that makes studying the diffuse ISM of nearby galaxies challenging.
Cosmological background emission
is a T = 2.73 K black body (cf. Sect. 1.2.4.2; Mather et al., 1994). It contaminates essentially the mm regime. Its emission is globally isotropic with some fluctuations. The amplitude of these fluctuations is shown as a solid red line in Fig. 3.24. These fluctuations are a bit more difficult to subtract. However, in general this source of contamination is not the most challenging to remove.
Non-dusty contamination,
such as gas-phase line emission or the radio continuum need to be subtracted. For instance, in the submm regime, the CO lines and the free-free continuum can account for 10 20% of the emission around λ 1 mm (cf. Fig. 2.10 and e.g.Galliano et al., 2003, 2005). Additional independent constraints are necessary, such as spectroscopic observations of the contaminating lines, and long-wavelength radio continuum fluxes that probe the free-free and synchrotron emission.
Differences in angular resolution
are due to variations of the beam size across the observed SED. If not corrected, this can have dramatic consequences, especially if the studied region has large gradients of emission on scales of the order of the largest beam size. It is however easy to correct. One needs to identify, among the instruments used, which one has the largest beam, Ωmax. One then simply needs to degrade the angular resolution of all the other wavelengths, to the resolution Ωmax. This degradation is performed by convolving the images by a kernel, which is the Ωmax beam deconvolved by the beam at the nominal wavelength. Such kernels are provided, for instance, by Aniano et al. (2011) for a wide variety of instruments.
Differences in field of view
come from the fact that the different instruments do not necessarily have the same orientation on the sky and the same pixel size. To model a spatially-resolved SED, it is thus necessary to reproject every image on a single grid. Numerous methods are available to perform this reprojection (e.g. Bertin et al., 2002). It can become problematic only when there are missing areas, such as incomplete maps or masked regions (e.g. because of saturation). The linear pixel size of the final grid can be as low as 13 of the largest beam size (Nyquist sampling).
Uncertainties
must be taken into account as rigorously as possible. Ideally, we should not only determine the uncertainties of each pixel at each wavelength, but also their correlations. This can be simplified by separating the two major sources of uncertainties: noise and calibration effects. The noise comes from the instrument. It is usually provided by the data reduction pipeline. It thus needs to be propagated through the various sources of homogenization we have discussed. In particular, resolution degradation increases the median signal-to-noise ratio, as it has an averaging effect on the noise spatial distribution. It also creates correlations between pixels. To our knowledge, the best method to propagate noise uncertainties, which is also the simplest to implement, is by way of Monte-Carlo simulations (e.g. Galliano et al., 2011). The principle is the following.
1.
We generate a large number, N 100, of images with flux:
Fν(i)(x,y,λ) = F ν(x,y,λ) + δ(i,x,y,λ) × σν(x,y,λ) (3.41)

(i = 1,N), where Fν and σν are the flux and uncertainty coming from the data reduction pipeline. The random variable, δ, is independent, normal with mean 0 and standard deviation 1.

2.
We then perform the different homogenization steps (contamination subtraction, degradation, reprojection) on each random samples.
3.
We now have, for each pixel and each wavelength of the final homogenized maps, a distribution of N values. The standard deviation of this distribution gives the uncertainty and we can compute correlation coefficients between different pixels or wavelengths.

Calibration uncertainties can be computed afterwards, as they are proportional to the flux. These uncertainties are fully correlated between pixels and partially correlated between wavelengths (e.g. Galliano et al., 2021).

This technical data preparation can be tedious, but it is crucial as dust model results directly rely on it. A significant effort has been put into providing homogenized databases of nearby galaxies. Among them, the most important surveys are the following.

PIC
Figure 3.24: Principal sources of contaminations encountered when modeling SEDs. The zodiacal spectrum (blue) has been computed using the model of Fixsen & Dwek (2002). The CMB spectrum is a T = 2.73 K black body (red dashed line). The solid red line corresponds to the level of CMB fluctuations. The CIB (green) comes from the compilation of Dole et al. (2006). The Galactic cirrus emission is the THEMIS model scaled by a column density NH = 1024 H/m2. Licensed under CC BY-SA 4.0.

Properties of individual galaxies. Numerous studies have presented the SED modeling of nearby galaxies, and their derived dust properties, either globally or spatially-resolved. We have participated to several such projects (e.g. Whaley et al., 2009; Galametz et al., 2009; Boselli et al., 2010; Galametz et al., 2010; O’Halloran et al., 2010; Eales et al., 2010; Cortese et al., 2010; Sauvage et al., 2010; Bendo et al., 2010; Roussel et al., 2010; Gordon et al., 2010; Davies et al., 2010; Boquien et al., 2010; Skibba et al., 2012; De Looze et al., 2012a; Galametz et al., 2013; Ciesla et al., 2014; Gordon et al., 2014; Galametz et al., 2016; Bianchi et al., 2018; Nersesian et al., 2019). Reviewing them would be unwieldy, here. In general, these studies provide dust parameters (mass, starlight intensity, PAH mass fraction, etc.) of different objects, which can be used in combination with other tracers to refine our understanding of the studied source. They also provide scaling relations and calibrations of various diagnostics such as SFR tracers. These results can also be used to train machine-learning models that could predict the SED of a poorly-observed galaxy (e.g. Dobbels et al., 2020).

Identifying dust heating sources. A particular question, that has been tackled by several studies, is the identification of the sources responsible for dust heating within galaxies. In the MW, 3D reconstruction of the ISM distribution showed that the heating by young, O/B stars (Table 3.2) was prominent in molecular regions, whereas the atomic phase was mainly heated by lower-mass stars (e.g. Sodroski et al., 1997; Paladini et al., 2007). In nearby galaxies, this depends on the SF activity of the galaxy. For instance, we showed that PAHs were essentially heated by field stars in the quiescent galaxy NGC 2403. These molecules are however heated by the escaping radiation from H II regions in the more actively star-forming object, M 83 (Jones et al., 2015). More generally, with the DustPedia sample, we found that dust in ETGs was mainly heated by old stars (Nersesian et al., 2019). It is only when considering more gas-rich galaxies that the contribution of young stars becomes more important. It can account for up to 60% of the dust luminosity in extreme late-type galaxies (Sm–Irr, Fig. 3.22;  Nersesian et al., 2019). These different heating sources have an impact on the global escape fraction (i.e. the fraction of stellar radiation leaving the galaxy unattenuated). Massive stars being embedded in molecular cocoons, they have a lower escape fraction than Low- and Intermediate-Mass Stars (LIMS) which occupy lower density regions. In the DustPedia sample, we showed that the escape fraction was on average 81%, with mild variations across galaxy types (Bianchi et al., 2018). It is slightly lower in LTGs ( 75%). We emphasize that this nearby galaxy sample lacks the deeply enshrouded star formation of LIRGs and ULIRGs, where the global escape fraction can drop down to 1% (e.g. Clements et al., 1996). The question of the dust heating contribution can now be tackled with more accuracy using 3D MCRT models.

Large-scale radiative transfer models of galaxies. Applying a 3D MCRT model to reproduce the spatial flux distribution of galaxies, in all wavebands, is not straightforward. Indeed, the observations provide only 2D projected constraints. This is why most studies favor edge-on galaxies, as the images of such objects provide constraints on both the radial and azimuthal distributions, assuming axisymmetry (Fig. 3.25). Several studies have modeled the effect of extinction on the optical data of disk galaxies using such codes (e.g. Xilouris et al., 1999; Alton et al., 2004; Bianchi, 2007). They were able to answer the recurring question about the optical thickness of disk galaxies (Disney et al., 1989). In particular, Xilouris et al. (1999) found that the face-on optical depth of typical spiral galaxies is less than one, in all optical bands. Concerning dust heating, recent progress has been made, especially by the DustPedia collaboration, using the MCRT code SKIRT (Baes & Camps, 2015).

A model such as SKIRT can also be used to model the radiative transfer in simulations of galaxies (e.g. Trčka et al., 2020). Finally, these models account for the energy balance between the escaping UV-visible light and the re-emitted IR-submm radiation. Interestingly enough, several studies report a deficit of modeled FIR emission by a factor 3 4, compared to the observations (Alton et al., 2000, 2004; Dasyra et al., 2005; De Looze et al., 2012a,b). This discrepancy is thought to result from a lack of detail in modeling the geometry. In particular, the presence of young stars, deeply embedded in molecular clouds, at sub-grid resolutions, could compensate for this deficit without significantly altering the extinction (cf. e.g. Baes et al., 2010).

PIC
Figure 3.25: Radiative transfer modeling of NGC 4565. The observations (left column) are compared to the modeled flux distribution (right column). Credit: De Looze et al. (2012a).

3.1.3.3 Constraining the Grain Opacity

Dust masses derived from SED fits directly depend on the assumed grain opacity. Using the MBB parametrization (Eq. 1.72), both the scaling, κ0, and the emissivity index, β, are important. There are particular situations, where we can reverse the process and constrain these two parameters:

1.
if we are observing a region that we can assume uniformly illuminated, then we can infer β;
2.
if we have an independent constraint on the dust mass, then we can infer κ0.

Studies of the emissivity index. There are numerous publications presenting MBB fits of nearby galaxies. However, as discussed in Sect. 3.1.2.1, the derived emissivity index, β, is degenerate with temperature mixing. The best constraints on the intrinsic β are obtained in the submm regime, where only massive amounts of very cold dust (T 10 K) could bias the value. Table 3.3 lists effective emissivity indices, βeff, for several objects, obtained with Planck, with constraints up to 850μm. It appears that all the values are lower than 2, and that low-metallicity systems have a lower βeff than higher metallicity galaxies. Boselli et al. (2012), studying a volume-limited sample with Herschel (up to 500μm), also found an average βeff 1.5, and hinted that low-metallicity objects tend to have βeff < 1.5. In M 33, βeff derived from Herschel observations is around 2 in the center and decreases down to 1.3 in the outer parts (Tabatabaei et al., 2014). On the other hand, the outer regions of M 31 exhibit a steeper slope (βeff 2.3) than in its center (Draine et al., 2014). This contradictory behaviour does not appear to originate in fit biases, as both increasing and decreasing trends of βeff with radius are found in the sample of Hunt et al. (2015).






Milky Way M 31 LMC SMC





Temperature

19.7 ± 1.4 K 18.2 ± 1.0 K 21.0 ± 1.9 K 22.3 ± 2.3 K





β eff

1.62 ± 0.10 1.62 ± 0.11 1.48 ± 0.25 1.21 ± 0.27





Reference

Planck (2014a) Planck (2015a) Planck (2011a) Planck (2011a)





Table 3.3: Free emissivity index MBB fits of nearby galaxies by the Planck collaboration.

Grain opacity in the LMC. An important result, that has often been misunderstood, concerns the grain opacity in the LMC. Galliano et al. (2011), modeling a strip covering one fourth of the LMC (cf. Fig. 3.21), with the composite approach (Sect. 3.1.2.2), found that the dustiness distribution in this galaxy was most of the time larger than the maximum value it could in principle reach. This maximum value is set by the elemental abundances. The fraction of elements locked-up in grains can indeed not be larger than the amount available in the ISM. The metallicity of the LMC is ZLMC 12Z (Pagel, 2003). The maximum dustiness of the LMC is thus Zdustmax Z LMC 0.007. The dust mixture that was used is an update of the Zubko et al. (2004, BARE-GR-S). It is essentially based on the Draine & Li (2007) optical properties, a pre-Herschel model. This discrepancy is shown in Fig. 3.26 (the red histogram). The only explanation is that this grain mixture is not emissive enough to account for the observed FIR-submm emission. We thus proposed an alternate dust model, simply replacing graphite by amorphous carbons (the ACAR sample of  Zubko et al., 1996), without altering the total carbon fraction. This simple modification boosts the emissivity by a factor 6 2 3. With this new model, most of the dustiness distribution is centered around its expected value, and is clear from the forbidden range (yellow). It was called the AC model (blue histogram in Fig. 3.26). The tail of the distribution in the forbidden zone originates in cold regions, where the uncertainty is large. The conclusion of this modeling was that LMC grains must be a factor 2 3 more emissive than the Draine & Li (2007) model. We actually presented a preliminary version of this result during Herschel’s science demonstration phase (Meixner et al., 2010).

PIC

Figure 3.26: Dust mass discrepancy in the LMC. The histograms are the pixel distribution of the dustiness in the LMC strip of Galliano et al. (2011). We show two models: (i) the standard model” (red), based on the Draine & Li (2007) optical properties; (ii) the “AC model (blue), replacing graphite by amorphous carbon to boost the emissivity. Most pixels of the standard model are in the hatched area, thus violating the constraints put by heavy element abundances. We also show the uncertainties on the limit dustiness (yellow error bar) and on the expected one (green error bar). Licensed under CC BY-SA 4.0.

Confirmation in other systems. The conclusion of the Galliano et al. (2011) study was that grains in the LMC had to be more emissive than in the MW, because the Draine & Li (2007) model was at the time consistent with the MW. This last statement however happened to be inexact. Planck Collaboration et al. (2016c) modeled the all sky dust emission, using also the Draine & Li (2007) model. The A(V ) estimated along the sightlines of 200000 Quasi-Stellar Objects (QSO) was systematically lower than their dust-emission-derived A(V ). Their comparison of emission and extinction thus indicates that the Galactic opacity is in fact also a factor of 2 higher than previously assumed. In addition, in M 31, Dalcanton et al. (2015) derived a high spatial resolution map of A(V ). As in the Galaxy, the emission-derived A(V ) map (Draine et al., 2014) was found to be a factor of 2.5 higher. We emphasize that, although each of these studies found evidence of local variations of the emissivity as a function of the density (cf. Sect. 4.2.1.1), the overall opacity seems to be scaled up compared to Draine & Li (2007). In other words, in all the environments where enough data is available to constrain κ, it is found a factor of 2 3 higher than the original Draine & Li (2007) properties. Dust models therefore need to use an opacity consistent with these constraints. This is the case of the THEMIS model. Its FIR-submm opacity is very close to our AC model (cf. Fig. 4 of Galliano et al., 2018).

 It is reasonable to adopt the THEMIS grain opacity (cf. Sect. 2.3.3), when modeling galaxies.

The opacity in nearby galaxies. The DustPedia collaboration conducted several studies aimed at constraining the grain opacity in nearby galaxies. First, Bianchi et al. (2019) studied the actual emissivity of 204 late-type galaxies, that is the ratio of IR emission to H column density. We found an emissivity 𝜖ν(250μm) 0.82 ± 0.07 MJy/sr/(1020 H/cm2), consistent with the MW, except for the hottest sources. These estimates were derived using global fluxes, integrated over the whole galaxy. In parallel, Clark et al. (2019) modeled in details the two face-on galaxies, M 74 and M 83. We could map the grain opacity. This was done by converting metallicity maps into oxygen depletion maps, and comparing those to the dust mass. The derived opacities were quoted at λ = 500μm: κ(500μm) 0.11 0.25 m2/kg in M 74, and κ(500μm) 0.15 0.80 m2/kg in M 83. These values are consistent with the Herschel-Planck-revised opacities of the THEMIS model (κ(500μm) 0.19 m2/kg; cf. Fig. 2.26).

3.1.3.4 Constraining the Size Distribution

We have seen in Sect. 3.1.2.2 that there is a degeneracy between the grain size and starlight intensity distributions. This degeneracy arises from the fact that it is observationally difficult to distinguish the MIR emission of a hot region from the MIR emission of small grains. In the early 2000s, we did not know it was impossible, so we did it (Galliano et al., 2003, 2005). We modeled the SED of the following four BCDs: NGC 1569 (Z 14Z), II Zw 40 (Z 16Z), He 2-10 (Z 1Z)and NGC 1140 (Z 13Z), to infer their size distribution. We interpreted these results in light of shock processing.

Grain processing by shock waves. Shock waves from SNe process dust grains, while sweeping the ISM.

Fig. 3.27 shows the model of Jones et al. (1996) for graphite and silicates. In both panels, the grey curve is the initial, MRN size distribution. The color curves show the size distribution obtained after the mixture has been swept by a shock of velocity, vshock. The main effect of the blast wave is to fragment and shatter grains, turning large grains into smaller grains. The qualitative effect is similar for both compositions. This is best seen at low velocity (blue curves). Large grains are depleted and there is an excess of small grains. At higher velocity, the distribution tends toward a log-normal centered around a 10 nm (magenta curve). Vaporization also leads to a net loss of dust mass. This model has since then been refined by Bocchio et al. (2014), who applied it to THEMIS grains. We will discuss more extensively dust processing by SN blast waves in Chap. 4.

PIC
Figure 3.27: Effect of a blast wave on the grain size distribution. In both panels, we demonstrate the processing of a MRN size distribution (grey) by shock waves of different velocities, vshock. Panel (a) shows graphite, and panel (b), silicates, both from the Jones et al. (1996) model. Licensed under CC BY-SA 4.0.

The modeling strategy. Galliano et al. (2003, 2005) modeled the UV-to-mm global SED of the four BCDs, using the DBP90 dust model, the stellar evolutionary synthesis code PÉGASE (Fioc & Rocca-Volmerange, 1997, Fig. 3.19), and the photoionization code, CLOUDY (Ferland et al., 1998). The modeling scheme was the following.

1.
We assume that the ISM is concentrated in a spherical shell. At the center of this shell are the stellar populations.
2.
We model the escaping UV-visible SED with two PÉGASE stellar populations, a young and an old one.
3.
We further constrain the UV-visible fit by selecting intrinsic stellar spectra that have the appropriate hardness. The hardness of the ISRF is constrained by matching the observed ISO [Ne III]15.56μm/[Ne II]12.81μm and [S IV]10.51μm/[Ne III]15.56μm ratios, using CLOUDY.
4.
The intrinsic ISRF is then used to heat the DBP90 dust mixture. We fit the dust emission to the IR-to-mm observations, by varying the size distribution.
5.
We iterate this process a few times, as the dust size distribution impacts the extinction curve, which is circularly used to constrain the stellar populations.

This model is self-consistent per se. It however avoids the degeneracy between size and ISRF distributions by assuming a simple geometry (the shell). The dust is thus uniformly illuminated in this model, which is an unrealistic assumption for a star-forming galaxy. It is therefore likely that some of the emission we have attributed to small grains originates in hot regions. The results Galliano et al. (2003, 2005) obtained, that we will discuss in the following paragraphs, are nevertheless qualitatively consistent with the properties we expect in these environments. It is possible that only a fraction of the MIR emission originates in compact H II regions, not excluding an overabundance of small grains.

The grain size distribution in four dwarf galaxies. The inferred size distribution, in the four BCDs, are shown in Fig. 3.28. We display the three components of the DBP90 model: PAHs, VSGs and BGs (cf. Sect. 2.1.2.5). The most striking features, common to the four objects, are the following.

Two studies used a similar approach, assuming uniform illumination, and fitting the SED varying the size distribution of the DBP90 model, in NGC 1569 (Lisenfeld et al., 2002) and in the LMC (Paradis et al., 2009). They also concluded to an overabundance of small grains.

PIC
Figure 3.28: Grain size distribution in four dwarf galaxies. In each panel, we show the MW grain size distribution of the DBP90 model, as dashed lines. Those are the three components presented in Sect. 2.1.2.5: PAH, VSG and BG. The filled curves show the likely range of size distribution inferred by the SED fits of Galliano et al. (2003, 2005). Licensed under CC BY-SA 4.0.

Consequence on the extinction curves. We have briefly mentioned in Sect. 2.1.2.2 that the extinction curves in the Magellanic clouds were systematically different from the MW. Fig. 3.29.a compares the extinction towards different sightlines in the LMC and SMC, to the range of extinction curves in the MW. We see that the LMC (Z 12Z) is on average (red curve) similar to the MW. However, toward the massive star-forming region 30 Doradus (LMC2 supershell; green curve), it is steeper, with a weaker 2175 Å bump. When we go to the SMC (Z 15Z; blue curve), the difference is more pronounced: the extinction curve has a very steep UV-rise and lacks the 2175 Å bump. The origin of these variations are still debated. Our four BCDs brought an interesting perspective on this open question. The extinction curves derived from the size distributions of Fig. 3.28 are shown in Fig. 3.29.b. We can see that they are systematically steeper than the average MW (R(V ) = 3.1). They also have a 2175 Å that is either weaker (NGC 1569, He 2-10 and NGC 1140) or similar (II Zw 40) to the MW 7. In other words, these extinction curves lie between the LMC and SMC, consistent with the metallicity range of these BCDs. The modeling of Galliano et al. (2003, 2005) therefore provides a coherent view of the grain properties in these environments, where shock waves have an instrumental role in shaping the grain sizes.

PIC

Figure 3.29: Extinction curves of low-metallicity environments. In both panels, we show as a reference the range of Galactic extinction curves, with R(V ) = 2 5, from the Fitzpatrick et al. (2019) sample (cf. Fig. 2.11), in grey. The white curve is the average Galactic extinction, with R(V ) = 3.1. In panel (a), we overlay several extinction curves within the Magellanic clouds, from the Gordon et al. (2003) sample: (i) the average of the LMC (red); (ii) the LMC2 supershell (green), near the massive star-forming region 30 Doradus; and (iii) the SMC bar (blue). In panel (b), we show the extinction of the four BCDs modeled by Galliano et al. (2003, 2005). Licensed under CC BY-SA 4.0.

3.2 Studies Focussing on Specific Spectral Domains

Sect. 3.1 was devoted to modeling the whole IR SED, which is necessary to estimate the total dust content of galaxies. There are however several other important properties that can be self-consistently studied by focussing on a particular wavelength range.

3.2.1 Scrutinizing Mid-IR Spectra

Mid-IR spectra have been extensively observed since the first light of ISO. Spitzer and AKARI have extended our knowledge of this spectral range and the JWST will likely revolutionize it.

3.2.1.1 The Aromatic Feature Spectrum

Until now, we have discussed PAHs from a general point of view, and how to estimate their mass fraction. We now focus on the information that the analysis of their detailed MIR emission spectrum can bring. In what follows, we interchangeably use the terms UIBs, aromatic features and PAH bands. The only case where these terms are not equivalent is when discussing the UIBs around 3 μm, coming from a mixture of aromatic and aliphatic features, that can not be solely attributed to PAHs.

MIR spectra of galaxies. Fig. 3.30 illustrates the diversity of MIR spectra encountered in different environments. It shows three extreme galaxies from the Hu et al., in prep. sample.

Gas-rich, Solar-metallicity galaxies,
such as NGC 1097 (green spectrum in Fig. 3.30), have bright aromatic features. These features are emitted by both their diffuse ISM and their PDRs. The level and steepness of their MIR continuum, longward 10 μm, depend on their SFR. Galaxies with a low SF-activity tend to have a flatter continuum, as the emission from hot equilibrium grains in H II regions is lower.
Low-metallicity galaxies,
such as NGC 1569 (blue spectrum in Fig. 3.30), have an integrated spectrum very similar to an H II region (e.g. Peeters et al., 2002b; Martín-Hernández et al., 2002). They have weak or undetected aromatic features, which we will extensively discuss in Sect. 4.2.2.1. Their strong ionic lines result from the combination of their young stellar population and lower dust screening. For the same reason, their MIR continuum is significantly steeper than Solar-metallicity objects having the same specific Star Formation Rate (sSFR SFRM). In addition to small, stochastically-heated grains, the MIR continuum of BCDs originates partly in large grains at thermal equilibrium with the radiation field in H II regions. The latter grains are indeed hot enough to significantly emit at shorter wavelengths than Solar-metallicity objects.
AGNs,
such as CentaurusA (red spectrum in Fig. 3.30), have a rather flat NIR-to-MIR continuum (e.g. Laurent et al., 2000; Spoon et al., 2007). This continuum originates in the strong gradient of temperature due to the central illumination by the AGN (this is the same qualitative effect as in our MCRT simulation; Sect. 3.1.1.4). In particular, temperatures of the underlying continuum can reach higher values in the central region, than normal galaxies. It explains why the NIR continuum does not drop to zero. These central hot regions produce a bright MIR continuum, attenuated by the entire disk, resulting in deep silicate absorption bands.
PIC
Figure 3.30: MIR spectra of galaxies. We show the AKARI and Spitzer spectra of three galaxies: (i) CentaurusA (red) is an ETG with a powerful AGN; (ii) NGC 1097 (green) is a LTG with a low-luminosity AGN; and (iii) NGC 1569 (blue) is BCD that we have scaled down by a factor of 3. These spectra were prepared for Galliano et al. (2018). We have indicated most of the relevant features. Silicate absorption bands are shown in grey, and ice absorption bands in cyan (cf. Sect. 2.2.1.2). Gas-phase emission lines are indicated in orange, and carbon-dust emission features, in magenta. Licensed under CC BY-SA 4.0.

Laboratory and theoretical PAH physics. Although we do not know the exact composition of the interstellar carbon grain mixture responsible for the aromatic feature emission, the brightest bands have been attributed to the main vibrational modes of PAHs. There are still some debates about the origin of the weakest features (cf. e.g. Allamandola et al., 1999; Verstraete et al., 2001; Tielens, 2008; Boersma et al., 2010; Jones et al., 2013). In Fig. 3.30, we have labeled the different bands with a given mode. These modes are schematically represented in Fig. 3.31.a.

The charge
of the molecules is one of the most important parameters controlling the ratio between the C–H and C–C bands. Fig. 3.32.a shows the laboratory data of Allamandola et al. (1999). The two spectra are the sum of several neutral and cationic molecules. It is clear that C–C and C–C–C bands (6.2, 7.7 and 17 μm complexes) are predominantly carried by ionized PAHs, whereas C–H bands (3.3, 11.3 and 12.7 μm complexes) are carried essentially by neutral PAHs.
Dehydrogenation
has a similar effect to ionization. However, for PAHs with more than 25 C atoms (i.e. the bulk of interstellar PAHs), hydrogenation through reactions with abundant atomic H is more important than H loss through unimolecular dissociation (cf. e.g. Hony et al., 2001). Thus, dehydrogenation does not have a detectable effect on the UIB spectrum.
The molecular structure
is another factor. C–H Out-Of-Plane (OOP) bending modes have different frequencies, depending on the number of H atom per aromatic cycle (cf. Fig. 3.31.b). The 11.3 μm band corresponds to a solo H, found on straight molecular edges, whereas the 12.7 μm one corresponds to a trio, found on corners of the molecules. The solo-to-trio intensity ratio, I11.3I12.7, is thus an indicator of PAH compactness (Iλ being the integrated intensity of the feature centered at λμm).
The size
of the PAHs affects the relative intensity of the different bands. This is demonstrated in Fig. 3.32.b. Small PAHs (magenta and red spectra) fluctuate up to temperatures higher than large ones (blue and cyan spectra). Short-wavelength bands are therefore more pumped in small PAHs, whereas large PAHs emit predominantly long-wavelength features.
ISRF hardness
has an effect similar to the size, as a higher mean photon energy causes the grain to fluctuate up to higher temperatures (Sect. 3.2.1.3 and e.g. Galliano et al., 2008b). We have however seen, in Fig. 3.14, that this effect was probably less than a factor 2, in astrophysically relevant cases.

 The charge and size of the PAHs are the main parameters controlling their emission spectrum.

PIC
Figure 3.31: Vibrational modes of PAHs. Panel (a) represents the main in-plane and out-of-plane vibrational modes. Panel (b) gives examples of solo, duo, trio and quartet H sites. In both panels, red spheres represent C atoms, and cyan spheres, H atoms. Licensed under CC BY-SA 4.0.

PIC

Figure 3.32: Laboratory and theoretical PAH spectra. Panel (a) shows the absorption coefficient of neutral and cationic PAHs measured in the laboratory by Allamandola et al. (1999). Panel (b) shows theoretical emission spectra of neutral and cationic PAHs, with radii a = 0.35 nm and a = 5 nm. We have used the Draine & Li (2007) optical properties and computed the stochastic heating for the Solar neighborhood ISRF (U = 1). Licensed under CC BY-SA 4.0.
3.2.1.2 Spectral Decomposition Methods

To study variations of the aromatic feature spectrum with environmental conditions, one needs to measure the intensity of each band. This task is not as simple as it appears. There are indeed several challenges.

MIR spectral decomposition methods are therefore an essential tool to properly study PAH features.

Calibrating feature properties. Since the properties of the different UIBs are not a priori known, we need to empirically infer them. In our work, we parametrize the band spectral profile with a split-Lorentz function (Hu et al., in prep.):

Fν = I× 2 π Δνs2 Δνs + Δνl 1 (ν ν0)2 + Δνs2  for ν ν0 2 π Δνl2 Δνs + Δνl 1 (ν ν0)2 + Δνl2  for ν < ν0, (3.42)

where ν0 is the central frequency of the feature, and Δνs and Δνl are its widths on the short- and long-wavelength sides. Having an asymmetric feature is indeed necessary to accurately fit good quality spectra. This asymmetry may originate in the anharmonicity of the transition responsible for the band, or may be due to unresolved blended features. The parameters characterizing each individual features, ν0, Δνs and Δνl, could be derived from each fit. However, most of them would be quite uncertain, using an average Spitzer spectrum. For that reason, we have calibrated these parameters (i.e. inferred their reference value), using high-resolution, high-signal-to-noise-ratio spectra of Galactic regions. This calibration is demonstrated in Fig. 3.33.

1.
We have first fitted the ISO/SWS spectrum of the reflection nebula, the Red Rectangle, in order to calibrate the narrow bands (cf. Fig. 3.33.a). The very wide bands, also called plateaus, are not very prominent in this region.
2.
We then fix the narrow band parameters and infer the λ 8μm and λ 12μm plateau parameters by fitting the ISO/SWS spectrum of the planetary nebula, NGC 7027 (cf. Fig. 3.33.b). The λ 17μm complex is too weak to be calibrated using this spectrum.
3.
The λ 17μm complex is calibrated by fitting the IRS spectrum of the PDR, M 17 (cf. Fig. 3.33.c).
4.
The aromatic and aliphatic bands in the λ 3μm region are calibrated by fitting the λ = 3 5μm spectrum of NGC 7027 (cf. Fig. 3.33.d). The continuum in this range is indeed clean and there are no other blended bands.

Table 3.4 gives the resulting band parameters. With these parameters fixed, we can fit even low-signal-to-noise-ratio spectra varying only the intensity of each band (parameter I in Eq. 3.42).

PIC PIC
Figure 3.33: Empirical calibration of UIB profiles. In each panel, the black dots with error bars (barely visible) represent the observations. These are ISO/SWS spectra except for M 17, which is a Spitzer/IRS spectrum. The fitted model is the sum of the different individual components: (i) grey bodies for the dust continuum; (ii) Gaussian profiles for gas lines; and (iii) split-Lorentz profiles (Eq. 3.42) for UIBs. Licensed under CC BY-SA 4.0.





λ 0

Δλ s

Δλ l

Type





3.291 μm

0.020 μm

0.019 μm

Main





3.399 μm

0.011 μm

0.024 μm

Main





3.499 μm

0.077 μm

0.071 μm

Small





5.239 μm

0.025 μm

0.058 μm

Small





5.644 μm

0.040 μm

0.080 μm

Small





5.749 μm

0.040 μm

0.080 μm

Small





6.011 μm

0.040 μm

0.067 μm

Small





6.203 μm

0.031 μm

0.060 μm

Main





6.267 μm

0.037 μm

0.116 μm

Main





6.627 μm

0.120 μm

0.120 μm

Small





6.855 μm

0.080 μm

0.080 μm

Small





7.079 μm

0.080 μm

0.080 μm

Small





7.600 μm

0.480 μm

0.502 μm

Plateau





7.617 μm

0.119 μm

0.145 μm

Main





7.870 μm

0.170 μm

0.245 μm

Main





8.362 μm

0.016 μm

0.016 μm

Small





8.620 μm

0.183 μm

0.133 μm

Main





9.525 μm

0.107 μm

0.600 μm

Small





10.707 μm

0.100 μm

0.100 μm

Small





11.038 μm

0.027 μm

0.073 μm

Small





11.238 μm

0.053 μm

0.153 μm

Main





11.400 μm

0.720 μm

0.637 μm

Plateau





11.796 μm

0.021 μm

0.021 μm

Small





11.950 μm

0.080 μm

0.222 μm

Small





12.627 μm

0.200 μm

0.095 μm

Main





12.761 μm

0.081 μm

0.140 μm

Main





13.559 μm

0.160 μm

0.161 μm

Small





14.257 μm

0.152 μm

0.059 μm

Small





15.893 μm

0.178 μm

0.200 μm

Small





16.483 μm

0.100 μm

0.059 μm

Small





17.083 μm

0.496 μm

0.562 μm

Plateau





17.428 μm

0.100 μm

0.100 μm

Small





17.771 μm

0.031 μm

0.075 μm

Small





18.925 μm

0.037 μm

0.116 μm

Small





Table 3.4: UIB profile parameters. These are the parameters of Eq. (3.42). We have converted the frequencies in wavelengths: ν0 cλ0, Δνs c(λ0 Δλs) cλ0, Δνl cλ0 c(λ0 + Δλl). These are the parameters used for the work of Hu et al. (in prep.).

Fitting every feature at once. The total MIR spectrum contains the emission from different physical processes that have to be separated in order to accurately measure UIB intensities (cf. Fig. 3.30). Several models have tackled this problem since the ISO days (e.g. Verstraete et al., 1996; Boulanger et al., 1998; Laurent et al., 2000; Madden et al., 2006; Smith et al., 2007; Galliano et al., 2008b; Mori et al., 2012; Lai et al., 2020).

Individual UIBs
can be fitted with Lorentz profiles (e.g. Galliano et al., 2008b), Drude profiles (e.g. Smith et al., 2007) or split-Lorentz profiles (cf. Table 3.4 and Fig. 3.33; e.g. Hu et al., in prep.).
The dust continuum
can be fitted with a linear combination of grey bodies. The opacity of these grey bodies can be a simple MBB power-law (e.g. Smith et al., 2007) or the opacity of realistic materials (e.g. Hu et al., in prep.). This dust continuum likely originates in a combination of small, stochastically-heated grains, and large, hot equilibrium grains. The fitted parameters (mass and temperature) of each grey body therefore have to be considered as nuisance variables. They can not be interpreted physically, because of the uncertainty of their heating mechanism. They are employed only to decompose the continuum and UIB emissions.
Gas lines
can be fitted with Gaussian profiles. Table 3.5 lists the most prominent MIR gas lines and their central wavelengths. The width of the profile is determined by the spectral resolution of the instrument, except at very high spectral resolution (R λΔλ 30000), where the true width of the line can be resolved.
The stellar continuum
only contributes at short wavelength, and can be modeled with a Rayleigh-Jeans law.
The extinction,
which is essentially absorption at these wavelengths, can be modeled with a general extinction curve (cf. e.g. Fig. 2.24), assuming a simple geometry (cf. e.g. Sect. 3.1.1.2). The most prominent dust features are the two silicate bands at 9.8 and 18 μm (cf. Fig. 3.30). Ice absorption can also be non negligible. They can be taken into account the same way, as icy band profiles are well-constrained (cf. Fig. 2.12.b). The optical depth ratio between different ice species and the silicates however varies from one source to the other (e.g. Yamagishi et al., 2015). The optical depth therefore has to be independently estimated for each species (i.e. one needs to derive τdust, τH2O, τCO, etc.).

Fig. 3.34.a demonstrates such a fitting method on the total spectrum of M 82 (Galliano et al., 2008b). It is labeled as the Lorentzian method, as the UIBs are modeled with Lorentz profiles. This earlier model did not use all the bands given in Table 3.4, that we are now using.

PIC
Figure 3.34: MIR spectral fitting methods. Demonstration of the three MIR fitting methods, described in this manuscript, on the total ISO/CAM spectrum of the starburst galaxy, M 82 (Galliano et al., 2008b). The observations (the black error bars) are identical in each panel. The models of panels (a) and (c) are least-squares fits, whereas panel (b) is a spline interpolation. Panels (a) and (b) are from Galliano et al. (2008b). The fit of panel (c) has been performed only up to λ = 14μm, using PAHtat (Pilleri et al., 2012). In addition to the four components of Fig. 3.35, a continuum and the main gas lines are simultaneously fitted. Licensed under CC BY-SA 4.0.





Central wavelength

Species

Transition

Type





4.052 μm

H I

Brackett α

recombination





5.511 μm

H2

0–0 S(7)

ro-vibrational





5.908 μm

H I

Humphreys γ

recombination





6.109 μm

H2

0–0 S(6)

ro-vibrational





6.910 μm

H2

0–0 S(5)

ro-vibrational





6.985 μm

Ar II

2P32–2P12

forbidden





7.460 μm

H I

Pfund α

recombination





7.502 μm

H I

Humphreys β

recombination





8.025 μm

H2

0–0 S(4)

ro-vibrational





8.991 μm

Ar III

3P2–3P1

forbidden





9.665 μm

H2

0–0 S(3)

ro-vibrational





10.511 μm

S IV

2P32–2P12

forbidden





12.279 μm

H2

0–0 S(2)

ro-vibrational





12.369 μm

H I

Humphreys α

recombination





12.814 μm

Ne II

2P32–2P12

forbidden





15.555 μm

Ne III

3P2–3P1

forbidden





17.035 μm

H2

0–0 S(1)

ro-vibrational





18.713 μm

S III

3P2–3P1

forbidden





21.829 μm

Ar III

3P1-3P0

forbidden





25.890 μm

O IV

2P32–2P12

forbidden





28.219 μm

H2

0–0 S(0)

ro-vibrational





33.481 μm

S III

3P1–3P0

forbidden





34.815 μm

Si II

2P32–2P12

forbidden





35.349 μm

Fe II

6D52–6D72

forbidden





36.014 μm

Ne III

3P1–3P0

forbidden





Table 3.5: Most prominent MIR gas lines.

Alternative methods. Several other MIR spectral fitting methods have been dicussed in the literature. The two following ones are worth mentioning.

The spline method
consists in interpolating the spectrum under the main band complexes, using spline functions (e.g. Vermeij et al., 2002; Galliano et al., 2008b). It is demonstrated in Fig. 3.34.b. A first spline interpolation defines the continuum and a second one defines the band plateaus. The areas between the two continua (magenta filled) represent the plateaus in Fig. 3.34.b. The bands are then simply the intensity above the plateaus (cyan filled). The advantages of this method are: (i) it is simple to implement; (ii) it is very fast to run; and (iii) it does not suffer from the degeneracies between blended features, or between band wings and the continuum. It has however several limitations: (i) the choice of the spline anchor points is arbitrary, which results in systematic differences with other methods; (ii) at medium-spectral resolution (typical of ISO/CAM or Spitzer/IRS SL-LL modes) it is impossible to deblend features and lines, such as the 12.7 μm UIB and the [Ne II]12.81μm line; and (iii) this method does not provide meaningful uncertainty estimates. Galliano et al. (2008b) performed a systematic comparison of the Lorentzian and spline methods, and found that, although band intensities were different, the trends between band intensities or band ratios were consistent with the two methods.
The template method
consists in fitting a small number of synthetic spectra characteristics of different regions or different species. We have demonstrated this method in Fig. 3.34.c using PAHtat (Pilleri et al., 2012). This method uses four main components. These components were extracted from a sample of Galactic PDRs and PNe, using blind-signal separation methods (Berné et al., 2007; Joblin et al., 2008). In that sense, they are empirical, synthetic small grain spectra. These components are the following.
Neutral PAHs
(PAH0) are shown in Fig. 3.35.a. Their brightest C–C band is centered at λ 7.65μm.
Ionized PAHs
(PAH+) are shown in Fig. 3.35.b. They have similar band centers as PAH0, but different intensity ratios, as we have previously seen (cf. Fig. 3.32). Both PAH0 and PAH+ have similar spectral characteristics as class A spectra of Peeters et al. (2002a), found in H II regions, and thought to be “processed PAHs.
Large ionized PAHs
(PAHx), shown in Fig. 3.35.c, have a peculiar redshifted C–C band at λ 7.9μm, observed in PNe by Joblin et al. (2008). They are assumed to be 100 C atom grains. Those are reminiscent of the class B spectra of Peeters et al. (2002a).
Evaporating VSGs
(eVSGs), shown in Fig. 3.35.d, are thought to be PAH clusters of 500 C atoms. Their spectra have the characteristics of class C of Peeters et al. (2002a) seen in post-AGB stars, with a broad 7.9 μm band and no 8.6 μm feature. Peeters et al. (2002a) conjectured these could be “freshly-formed” carbon grains. The carrier of the broad 7.9 μm band could be destroyed by the ionizing radiation, during its journey in the ISM (e.g. Joblin et al., 2008). In this scenario, PAHs are formed by the destruction of VSGs at the UV-illuminated edges of molecular clouds. Class C is not well studied. Some novae show spectra similar to Class C, and could be due to nitrogen impurities in hydrocarbon compounds (Endo et al., 2021).

The advantages of this method are that: (i) fits are not extremely degenerate, even at low signal-to-noise ratios, because of the small number of free parameters, compared to the Lorentzian method; (ii) the four different classes are physically meaningful, providing a clear interpretation of the results. However, its lack of flexibility prevents accurate fits, that could overlook new information present in the observations.

PIC
Figure 3.35: PAH and small grain templates. We show the four components of PAHtat (Pilleri et al., 2012). Vertical yellow dashed lines indicate the same wavelengths in the four panels, helping visualizing the band shifts of some components. Licensed under CC BY-SA 4.0.

3.2.1.3 PAH Band Ratio Studies

Applying a spectral decomposition method to a set of MIR spectra allows us to study the variations of several band ratios that contain physical information about the small carbon grain properties.

Observed band ratios in galaxies. Fig. 3.36 demonstrates the diversity of spectra among galaxies (panel a) or within one (panel b). This figure emphasizes the differences in terms of aromatic band intensity. All these spectra are normalized by the 11.3-μm feature intensity. Yet, they exhibit large variations of their 6-to-9-μm features. This can be more precisely quantified by studying the correlation between specific band ratios, such as in Fig. 3.37 (Galliano et al., 2008b). The quantity I(λ) is simply the intensity of the feature centered at λμm. The bottom two panels show the results for integrated galaxies and Galactic regions, whereas the bottom two panels show the results for a few spatially-resolved sources. Overall the trends are similar for both types. It means that this is a multiscale relation, valid at sub-pc scales (within the Orion bar or M 17) and 10 kpc scales (among integrated galaxies). These three displayed band ratios span about an order of magnitude, and are linearly correlated with each other. It implies that the 6.2, 7.7 and 8.6 μm features are tied together, while the 11.3 μm can vary independently. This is what was illustrated in Fig. 3.36. The only parameter that can explain such a variation is the charge of the PAHs (cf. Fig. 3.32).

 The variation of the PAH charge can explain most of the UIB variations observed in the nearby Universe.

PIC
Figure 3.36: Diversity of MIR spectra among and within galaxies. Panel (a) shows the integrated spectra of three galaxies: (i) UM 448, a BCD; (ii) M 51, a prototypical LTG; and (iii)  NGC 4945, a LIRG. Panel (b) shows the spectra of three different regions within the starbursting irregular galaxy, M 82. In both panels, the monochromatic flux is normalized to its value at 11.3 μm, in order to demonstrate band ratio variations. These are ISO spectra (Galliano et al., 2008b). Licensed under CC BY-SA 4.0.

PIC

Figure 3.37: PAH band ratio correlations inside and among galaxies. The top two panels show the values of the ratios derived from integrated spectra of galaxies, as well as Galactic and Magellanic regions. The bottom two panels show pixel distribution of the same ratios, within a variety of objects. These data are a combination of ISO/CAM and Spitzer/IRS spectra, fitted using the Lorentzian method (Galliano et al., 2008b). The grey-filled curves represent the ± 1σ linear fit to the integrated source correlations (cf. Table 2 of Galliano et al., 2008b). The grey curves are identical in panels (a) and (c) and in panels (b) and (d), to guide the eye when comparing integrated and resolved correlations. Licensed under CC BY-SA 4.0.

Effects of ionization and size. Although ionization is the main driver of the UIB relative variations in galaxies, other effects can play a role in specific environments. The most important one of these secondary effects is the PAH size distribution. Fig. 3.38 shows numerical simulations of several key UIB ratios (Hu et al., in prep.). We have varied: (i) the minimum PAH size expressed in number of C atoms, NCmin, highlighted in Fig. 3.38.a; (ii) the PAH charge fraction, f+, highlighted in Fig. 3.38.b; (iii) the ISRF intensity and hardness. To illustrate the last effect, we have computed the model grid for the Solar neighborhood ISRF (Mathis et al., 1983). This values are the bright grid points. We have also computed the grid for a hot star spectrum, with U = 104 (the faint grid points). The effect of the ISRF is non negligible, but it is less drastic than that of the charge and size. Most astrophysically relevant ISRFs will be intermediate between the two extreme cases we have displayed. Fig. 3.38 illustrates the following points (see also  Rigopoulou et al., 2021, for a more complete calculation using Density Functional Theory).

I(3.3)I(11.3)
is mainly a size tracer. Both 3.3 and 11.3 μm bands (C–H modes) are indeed primarily emitted by neutral PAHs. Charge thus does not significantly impact this ratio. The 3.3 μm feature is mainly carried by the smallest PAHs, whereas the 11.3 μm feature is carried by intermediate sizes.
I(7.7)I(11.3),
on the opposite, is essentially a charge tracer. It is still sensitive to the size, because of the relatively large difference in wavelength of both features. The degeneracy due to this additional dependence can be broken, using I(3.3)I(11.3).

Such band ratio diagrams have been used to demonstrate systematic variations of the PAH size distribution in various environments. We studied the spatial variations of I(3.3)I(11.3) in NGC 1097 (Fig. 3.30) and showed it was systematically lower in the central region, close to the AGN (Wu et al., 2018b). The most likely explanation is that the hard radiation field from the central engine is selectively destroying the smallest PAHs. This effect was also shown by Smith et al. (2007) and Sales et al. (2010), on global scales.

PIC
Figure 3.38: Theoretical MIR band ratio variations. Both panels show the same grid points, but are highlighted differently. Each circle is the theoretical band ratio estimated from the stochastic emission spectrum of a PAH mixture having the Draine & Li (2007) optical properties and the Zubko et al. (2004, BARE-GR-S) size distribution. The solid color symbol are mixtures heated with the Solar neighborhood ISRF (Fig. 1.28) with U = 1, whereas the pastel symbols are heated by a black body at T = 30000 K with U = 104. In panel (a), we highlight the effect of changing the minimum size cut-off, in number of C atoms, NCmin. In panel (b), we show the effect of the charge fraction, f+. Licensed under CC BY-SA 4.0.

The particular case of low-metallicity environments. In low-metallicity systems, band ratio variations can be more difficult to probe, as the band equivalent widths are lower, and thus more uncertain (cf. Sect. 3.2.1.4). In the LMC, Mori et al. (2012) found different trends in neutral and ionized sightlines. Toward the latter, there are evidences that PAHs have a lower charge (as a consequence of the higher recombination rate) and are on average larger (due to the destruction of the smallest PAHs). In contrast, in the SMC, Sandstrom et al. (2012) found very weak I(6 9)I(11.3) ratios and weak 8.6 and 17 μm bands, implying small weakly ionized PAHs. This last point is consistent with the trend of I(17)I(11.3) with 12 + log (O/H) found by Smith et al. (2007). This was also noted by Galliano et al. (2008b), who found that low-metallicity systems tends to lie on average toward the lower left corner of Fig. 3.37, whereas the upper left corner is essentially populated by Solar-metallicity sources. However, Hunt et al. (2010) argued that BCDs exhibit a deficit of small PAHs. If there is a smooth variation of PAH size distribution with metallicity, these results are in contradiction. Sandstrom et al. (2012) noted that these BCDs are more extreme environments than the SMC, and that photodestruction could dominate the PAH processing (cf. Sect. 4.2.2.1). We note that the solution to this apparent controversy might alternatively reside in the difference in studied spatial scales. In the Magellanic Clouds, Spitzer spectroscopy gives a spatial resolution of a few parsecs, compared to a few hundred in nearby BCDs. The fact is that the LMC and SMC exhibit strong spatial variations of their UIB spectrum. Whelan et al. (2013) showed a diversity of MIR spectral properties in the SMC. In this study, we demonstrated that the PAH emission in a region like N66 is dominated by its diffuse component, and not by its bright clumps, where PAHs are destroyed. At the other extreme, the molecular cloud SMC-B1#1 shows unusually high UIB equivalent widths (Reach et al., 2000). Also, the I(11.2)I(12.7) ratio indicates that PAHs are more compact in 30 Doradus and more irregular outside (Vermeij et al., 2002). All these elements suggest that there is a complex balance of processes shaping the MIR spectra throughout low-metallicity environments.

UIBs as diagnostics of the physical conditions. The fact that ionization dominates the UIB variation in galaxies opens the possibility to use specific observed band ratios to quantify the physical conditions. The charge of an ensemble of molecules is indeed the balance between: (i) the ionizing photon rate; and (ii) the electronic recombination rate. The first quantity is usually quantified by the variable G0, defined as the integral of the ISRF in the FUV (e.g. Hollenbach & Tielens, 1997):

G0 6eV13.6eVI E(E)dE 1.6 × 106W/m2 . (3.43)

The recombination rate is roughly proportional to neT gas (ne being the electron density, and Tgas, the gas temperature; e.g.  de Jong, 1977). The ratio of these two rates, often called the photoionization parameter, therefore quantifies this equilibrium (e.g. Chap. 5 of Tielens, 2005):

γ G0 ne T gas . (3.44)

The electron density can be related to the total H density by considering that most electrons in the neutral gas come from the photoionization of C. We thus have ne (CH)nH 2 × 104n H (cf. Sect. 2.2.3.1). Galliano et al. (2008b) measured the I(6.2)I(11.3) ratio in Galactic regions where G0, nH and Tgas had been reliably estimated (Fig. 3.39.a). It allowed us to propose an empirical relation between γ and I(6.2)I(11.3):

I(6.2) I(11.3) G0(ne1cm3)T gas103K 1990 + 0.26 ± 0.16. (3.45)

In other words, measuring I(6.2)I(11.3) provides an estimate of γ. With such a relation, the diagrams of Fig. 3.37 can now be turned into the diagnostics of Fig. 3.40. This relation has, since then, been refined by several studies, especially Boersma et al. (2016). Although unique, the diagnostics of Eq. (3.45) has however the following limitations.

Finally, the calibration of Eq. (3.45) can also be estimated theoretically. Fig. 3.39.b shows the theoretical I(6.2)I(11.3) ratio as a function of G0 and ne (Galliano, 2009). It has been derived by computing the stochastic emission of PAHs within the PDR model of Kaufman et al. (2006). Such model indeed computes the charge balance of PAHs at each point within the cloud, where G0 and ne are known. One grid point value corresponds to a whole cloud.

PIC PIC
Figure 3.39: Calibration of PAH ratio diagnostics. Panel (a) shows the empirical correlation between the observed 6.2-to-11.3 μm ratio and the modeled parameter, γ (Eq. 3.44), in three galactic PDRs (Galliano et al., 2008b). Panel (b) shows the theoretical variation of the band ratio as a function of G0 and ne (Galliano, 2009). This parameter grid was computed implementing stochastic heating in the PDR model of Kaufman et al. (2006), using a fixed PAH size distribution (the BARE-GR-S of  Zubko et al., 2004). Licensed under CC BY-SA 4.0.

PIC

Figure 3.40: PAH band ratio as diagnostics of the physical conditions. These are the data shown in Fig. 3.37 where the I(6.2)I(11.3) ratio as been converted to G0ne ×T using Eq. (3.45) (Eq. (5) of  Galliano et al., 2008b). Licensed under CC BY-SA 4.0.

Other properties. We end this section by briefly reviewing other properties that can be studied with observations of UIBs.

PAH compactness
can be probed by studying the relative variations of C–H OOP bending modes (cf. Fig. 3.31.a). These bending modes depend on the number of H atoms per aromatic cycle (cf. Fig. 3.31.b). We have seen in Sect. 3.2.1.1 that, in particular, the solo-to-trio intensity ratio, I11.3I12.7, is an indicator of PAH compactness. This ratio was scrutinized in Galactic regions (evolved stars, H II regions, reflection nebulae and YSOs) by Hony et al. (2001). Their results are consistent with the scenario where large ( 100 150 C atoms), compact PAHs are formed in winds of evolved stars, and degraded into smaller, irregular molecules in the ISM.
Small a-C(:H) hydrogenation
can be studied with the I(3.4)I(3.3) ratio. There is a debate whether aromatic features are carried by PAHs or small a-C(:H). The 3.4 μm aliphatic feature however can not be carried by pure PAHs, it must come from either a-C(:H) grains (Jones et al., 2013) or Hydrogenated PAHs (HPAH) with one of several aliphatic groups (e.g. Fig. 1 of Marciniak et al., 2021). The I(3.4)I(3.3) aliphatic-to-aromatic ratio shows regional variations in the ISM, as the result of structural changes in the hydrocarbons through UV processing (e.g. Jones et al., 2013). Mori et al. (2014) showed that the I(3.4)I(3.3) decreases with the ionization of PAHs, in Galactic H II regions. Yamagishi et al. (2012) detected the 3.4 μm feature in the superwind of M 82. They found that the I(3.4)I(3.3) ratio increases with distance from the center. They interpreted this trend as the production of small a-C(:H), by shattering of larger grains in this harsh halo. Similarly, Kondo et al. (2012) found a higher I(3.4)I(3.3) ratio in the nuclear bar of NGC 1097, suggesting that the gas flow towards the center could lead to the formation of small a-C(:H) by shattering. We note that, alternatively, the I(3.4)I(3.3) ratio can increase with the accretion of a-C(:H) mantles in denser regions (Jones et al., 2013). This feature can also be seen in extinction, in AGNs (e.g. Mason et al., 2007). Hu et al. (in prep.) modeled the spatially resolved AKARI and Spitzer spectra in M 82. In this study, we found a negative correlation between the I(3.4)I(3.3) and [S IV]10.51μm/[Ne II]12.81μm ratios. The latter is a ISRF hardness indicator. This result thus demonstrates the dehydrogenation of a-C(:H) grains by hard ISRFs.
SFR indicators
are one of the most sought after astrophysical diagnostics. It happens that UIBs can be used to that purpose. Peeters et al. (2004) showed that the 6.2 μm feature intensity correlates well with SFR, making it a reliable estimator. Alternatively, Shipley et al. (2016) have calibrated a SFR estimator based on the integrated power of the 6.2, 7.7 and 11.3 μm features. The reason of this correlation is the same as for the TIR-SFR correlation (e.g. Kennicutt & Evans, 2012, for a review). At first order, the UIB strength is indeed correlated with TIR. Peeters et al. (2004) however note that UIBs are biased towards B stars. In addition, we note that such a tracer will be underestimating the SFR at low metallicity, because of the variation of the relative UIB strength, as we will see in Sect. 3.2.1.4.

3.2.1.4 Variations of the Aromatic Feature Strength

The evolution of the shape of the UIB spectrum, probed by studying band ratio variations, is not the only diagnostics of the small carbon grain properties. The overall aromatic feature strength, relative to the continuum (i.e. to the emission of the rest of the dust populations) shows drastic variations across environments (cf. Fig. 3.30). These variations trace the evolution of the mass fraction of their carriers – PAH or small a-C(:H).

Effect of ISRF hardness. PAH and small a-C(:H) are known to be sensitive to hard an intense radiation fields. They tend to evaporate near massive stars, and can be assumed to be fully depleted in H II regions (e.g. Cesarsky et al., 1996b; Galametz et al., 2013; Galliano et al., 2018). This effect can be quantified by studying the variation of the aromatic feature strength with a tracer of the ISRF hardness.

1.
The aromatic feature strength can be traced with I(PAH)/I(cont), the PAH-to-MIR-continuum ratio, where I(PAH) is the sum of the intensities of every aromatic feature, and I(cont) is the integrated intensity of the continuum, over an arbitrary wavelength range (10-16 μm in the case of  Madden et al., 2006).
2.
The hardness of the ISRF can conveniently be traced by the [Ne III]15.56μm/[Ne II]12.81μm ratio, as both of these MIR lines are usually bright. Madden et al. (2006) demonstrated, using a photoionization model, that [Ne III]15.56μm/[Ne II]12.81μm is around unity when the ISM is heated by a young star cluster, and drops rapidly after a few million years, as massive stars die.

Such a trend is shown on Fig. 3.41, with the Madden et al. (2006) results 8. It clearly indicates that PAHs are less abundant in regions permeated with a hard ISRF, at all spatial scales down to 0.1 pc. Note however that both tracers do not come from the same physical region: (i) PAH being destroyed in H II regions are tracing the neutral gas; (ii)  [Ne III]15.56μm and [Ne II]12.81μm obviously come from the ionized gas. The correlation therefore reflects the mixing of phases within the beam.

PIC
Figure 3.41: Effect of ISRF hardness on PAH strength. These results are from the spectral decomposition of Madden et al. (2006) and Galliano et al. (2008b). Panel (a) shows integrated sources, whereas panel (b) shows pixel-by-pixel distributions. The white line with the grey band shows the affine fit to the global values, ± 1σ, in log-log space. It is the same in both panels. Licensed under CC BY-SA 4.0.

Effect of metallicity. The effects of ISRF and metallicity are often degenerate, as at low metallicity: (i) stars of a given mass have a systematically higher effective temperature, because of line blanketing effects; (ii) the ISM is less opaque, because of the lower dustiness, and thus more permeated by UV radiation. The highest [Ne III]15.56μm/[Ne II]12.81μm ratios are found in BCDs, as well as the lowest I(PAH)/I(cont). Fig. 3.42.a shows the evolution of I(PAH)/I(cont) as a function of metallicity (Madden et al., 2006). There is a paucity of PAHs in low-metallicity environments. The questions is whether this is the result of their increased destruction, or if they have not been produced. We will come back to this question, when discussing dust evolution in Sect. 4.2.2.1.

 PAHs are under-abundant in low-metallicity environments.

The absence of metallicity threshold. The relation of Madden et al. (2006) was the first spectroscopy-based demonstration of the effect. Shortly before, Engelbracht et al. (2005) showed the broadband correlation of Fν(8μm)Fν(24μm) as a function of metallicity. They showed both quantities were clearly correlated. They however argued there were essentially two populations, below and above 12 + log (O/H) 8. Galliano et al. (2008a) showed that this was a bias due to the saturation of IRAC8μm as a PAH tracer at low metallicity. When the aromatic feature strength becomes indeed low, Fν(8μm)Fν(24μm) is not anymore a measure of I(PAH)/I(cont), but is a measure of the temperature of the continuum. This is illustrated on Fig. 3.42.b. We show that when the actual mass fraction of small a-C(:H) drops below 10%, Fν(8μm)Fν(24μm) becomes insensitive to its value.

PIC PIC
Figure 3.42: Effect of metallicity on the PAH strength. Panel (a) shows the trend of PAH relative strength with metallicity, by Madden et al. (2006). This was the first spectroscopy-based demonstration of this relation. Panel (b) illustrates the bias of the Fν(8μm)Fν(24μm) ratio as a tracer of PAH strength. It shows the flux ratio as a function of the modeled fraction of small a-C(:H) (the grains carrying the aromatic features), qAF, in the DustPedia sample (G21). Only galaxies with both IRAC8μm and MIPS24μm fluxes are displayed. Licensed under CC BY-SA 4.0.

3.2.2 Long-Wavelength Properties

At long wavelengths, in the submm-to-cm range, dust emission does not always behave as the extrapolation from the Rayleigh-Jeans law (Fν ν2+β). Two peculiar phenomena have been widely discussed in the literature: (i) the submillimeter excess; and (ii) the Anomalous Microwave Emission (AME).

3.2.2.1 The Elusive Submillimeter Excess

An excess emission above the modeled dust continuum is often observed, longward λ 500μm. The most significant reports of this submm excess can not be accounted for by: (i) thermal dust emission; (ii) free-free and synchrotron continua; and (iii) molecular line emission (cf. Fig. 3.43; Galliano et al., 2003).

1.
The first occurrence of such an excess was unveiled by Reach et al. (1995), studying the COBE observations of the Milky Way. Their IR–submm SED could be fitted with a MBB (β = 2; cf. Sect. 3.1.2.1), and an additional 4 7 K component.
2.
A few years later, Lisenfeld et al. (2002) and Galliano et al. (2003) found a statistically significant excess in the dwarf galaxy NGC 1569, at 850 μm and 1.3 mm. It is shown in Fig. 2.10.
3.
Several subsequent studies confirmed the presence of an excess in other late-type galaxies (e.g. Dumke et al., 2004; Bendo et al., 2006; Zhu et al., 2009; Galametz et al., 2009, 2011), including the global SEDs of the Magellanic clouds (Israel et al., 2010; Bot et al., 2010).
4.
Herschel and Planck opened the way to more detailed characterizations.

Studying this excess is important, as: (i) it could bias dust mass estimates; (ii) it potentially contains untapped physical information about the ISM.

PIC
Figure 3.43: Submillimeter excess in NGC 1569. This is the best fit model of Galliano et al. (2003), discussed in Sect. 3.1.3.4. We focus here on the blue component, which is a T = 5 K, β = 1 MBB accounting for 80% of the dust mass of the galaxy. This component has been fitted to account for the submm excess. Licensed under CC BY-SA 4.0.

Possible explanations. The origin of the submm excess is currently debated. The following explanations have been proposed. These different scenarios are not exclusive.

Very cold dust
(VCD), that is equilibrium grains colder than T 10 K, can be used to fit the excess. The emission of such a grain population indeed peaks in the submm range. Cold dust being weakly emissive, this scenario however leads to massive amounts of grains (40 80% of the total dust mass; e.g.  Galliano et al., 2003, 2005). Galliano et al. (2003) showed that VCD would be realistic only if this component was distributed in a few number of dense, parsec-size clumps. The existence of such cold dust indeed requires it to be shielded from all UV-visible light. Using the spatially resolved observations of the 500 μm excess in the LMC, Galliano et al. (2011) concluded that this explanation is unrealistic, as it would require at least one clump in each of the 90000 pixels of this study.
Temperature-dependent emissivity.
Laboratory measures on cosmic grain analogs exhibit an increase of the FIR-submm opacity as a function of temperature (e.g. Mennella et al., 1998; Demyk et al., 2017a,b). Meny et al. (2007) designed the so-called Disordered Charge Distribution/Two-Level System (DCD/TLS) model, predicting an increase of κ(λ0) and a decrease of β with the temperature of amorphous grains. It reproduces the submm excess observed in the Galaxy by Paradis et al. (2012) and in the LMC by Bot et al. (2010, coupled with spinning grains; cf. Sect. 3.2.2.2). However, this model can not account for the submm excess in the SMC (Bot et al., 2010).
Magnetic grains.
Draine & Hensley (2012) showed that the submm excess of the SMC could be attributed to magnetic nanoparticles (Fe, Fe3O4, γ-Fe2O3 9). Thermal fluctuations in the magnetization of these grains can produce strong magnetic dipole emission, since ferromagnetic materials are known to have large opacities at microwave frequencies. This hypothesis seems to be consistent with the observed elemental abundances of the SMC and could also be responsible for the excess detected in other environments. These magnetic nanoparticles could be free-flying or inclusions in larger grains.

Empirical properties of the excess. Since the origin of the submm excess is still unknown, most studies focus on characterizing its observed properties.

Low-metallicity systems
are the environments where the excess appears the be the most prominent. This is the reason why it has been detected in most nearby BCDs (e.g. Galliano et al., 2003; Dumke et al., 2004; Galliano et al., 2005; Galametz et al., 2009; Bot et al., 2010; Galametz et al., 2011). Galliano et al. (2011) have shown, using spatially-resolved observations of the LMC, that the SPIRE500μm excess, r500, varies up to 40% in certain regions. It is correlated with the mean starlight intensity, U, and anticorrelated with the dust surface density, as shown in Fig. 3.44. The correlation with U could be a confirmation of the DCD/TLS effect, as U directly controls the temperature of the grains. The correlation with surface density is however significantly better than with ISRF, a trend that the DCD/TLS model alone fails to explain.
In the Milky Way,
Paradis et al. (2012) showed that the SPIRE500μm excess becomes significant in the peripheral regions ( > 35), as well as towards some H II regions. This is qualitatively consistent with the trend we found in the LMC, as these peripheral regions are also the most diffuse ones. The relative amplitude of the excess can rise up to 20%.
In other galaxies,
the excess is more difficult to detect (e.g. Rémy-Ruyer et al., 2015; Dale et al., 2017). When resolved in non-barred spirals, the submm excess is primarily detected in the disk outskirts, thus at low-surface density (e.g. Hunt et al., 2015).

 The submm excess is more prominent in low-metallicity environments, and in diffuse regions.

PIC
Figure 3.44: Spatially-resolved submm excess in the LMC. In both panels, the color density represents the number of pixels per bin of parameters in the star-forming region N44 of the LMC (Galliano et al., 2011). The SPIRE500μm excess is shown as a function of: (a) mean starlight intensity (Eq. 3.38); and (b) dust mass surface density. Licensed under CC BY-SA 4.0.

Reality of the phenomenon. The reality of the submm excess has been questioned for the two following reasons. We try to address these criticisms in order to support its likeliness.

The excess is model-dependent.
Different dust opacities lead to different amplitudes of the excess. Some over-parametrized models have found excesses virtually everywhere. For instance, the Broken-Emissivity Modified Black Body (BEMBB) is a MBB with two free-varying emissivity indices, β1 and β2, below and above λ = λb (Gordon et al., 2014). With such a model, any excess can be fitted, but the derived β1, β2 and λb do not necessarily correspond to existing solids. This model is also extremely degenerate, as shown by Galliano (2018). A BEMBB fit of the diffuse Galactic ISM SED (cf. Fig. 2.25.d) by Gordon et al. (2014) gives a r500 48 ± 11% excess starting at λb 300 ± 30μm. Yet, this is the SED that everyone uses to calibrate dust models. It does not exhibit an actual excess. It illustrates that probing the submm excess with models which are not based on realistic optical properties is a non-sense. It must be studied with a model as physical as possible.
The technically-challenging nature of submm-mm observations
is also questioning the reality of the excess: (i) observations in this regime are difficult from the ground (cf. Sect. 2.1.1.1); (ii) the calibration of these observations is often uncertain; (iii) data reduction methods have problems dealing with the diffuse emission, which is where the excess appears to be the most prominent. In particular, Planck data have lead to revise the calibration of the COBE/FIRAS FIR-submm spectrum of the diffuse Galactic ISM (Planck Collaboration et al., 2014c). With this new calibration, the excess of Reach et al. (1995), discussed earlier, is not significant anymore. There are however several sound indications that these concerns are not enough to doubt the reality of the phenomenon.

3.2.2.2 The Anomalous Microwave Emission

As we have seen in Sect. 2.2.2.3, the AME is a centimeter continuum excess that can not be accounted for by: (i) the extrapolation of dust thermal emission; (ii) molecular line emission; and (iii) free-free and synchrotron continua (Fig. 2.10). Its SED looks like a bump peaking around λ = 1 cm (cf. Fig. 2.10). It is commonly attributed to the dipole emission of fastly rotating, small dust grains. The faster grains rotate, the shorter the frequency of the emission peak is.

The AME in extragalactic environments. In nearby galaxies, the first unambiguous detection of an AME has been obtained in an outer region of NGC 6946 (Murphy et al., 2010; Scaife et al., 2010). Follow up observations showed evidence for AME in eight regions of this galaxy (Hensley et al., 2015). This study showed that the spectral shape of this AME is consistent with spinning dust, but with a stronger AME-to-PAH-surface-density ratio, hinting that other grains could be the carriers. Overall, the AME fraction is highly variable, in nearby galaxies. Peel et al. (2011) put upper limits on the AME in M 82, NGC 253 and NGC 4945. These upper limits suggest that AME/100 μm is lower than in the Milky Way, in these objects. In M 31, Planck Collaboration et al. (2015a) report a 2.3σ measurement of the AME, consistent with the Galactic properties. Finally, Bot et al. (2010), fitting the NIR-to-radio SED of the LMC and SMC, tentatively explained the submm/mm excess with the help of spinning dust, in combination with a modified submm dust emissivity (cf. Sect. 3.2.2.1). They concluded that, if spinning grains are responsible for this excess, their emission must peak at 139 GHz (LMC) and 160 GHz (SMC), implying large ISRF intensities and densities. Draine & Hensley (2012) argued that such fastly rotating grains would need a PDR phase with a total luminosity more than two orders of magnitude brighter than the SMC.

Controversy about the carriers of the AME. Although PAHs have been considered the most likely carriers of the AME, Hensley & Draine (2017) argued that nanosilicates could be a viable alternative. They showed that nanosilicates can indeed account for the AME, without violating the other observational constraints (depletions, emission, extinction; cf. Sect. 2.2). This claim relies on the earlier findings of Hensley et al. (2016), showing that AME correlates better with dust radiance, R10, in the MW. Hensley et al. (2016) also found some fluctuations in the PAH emission relative to the AME intensity, traced by WISE12μm. The fact is that there is no observational evidence of nanosilicates in the ISM. In particular, these grains would emit the 9.8 and 18 μm features, as they would be stochastically heated. These bands would eventually remain diluted in the aromatic feature emission, provided the abundance of nanosilicates is low enough. We can however note that in H II regions, where PAHs are destroyed, we can see the underlying continuum, and silicate features in emission are rarely present (e.g. Peeters et al., 2002b). They can be seen only in the most compact H II regions, and their 9.8-to-18 μm ratio indicates that they originate in large equilibrium grains. A weak correlation does not always indicate an absence of causality. This issue might reside in the fact that Hensley et al. (2016) used the WISE12μm band as a tracer of PAH intensity, whereas this broad photometric band also encompasses a significant fraction of the underlying continuum, and is biased towards neutral PAHs. We have tried to address this controversy by refining the constraints on the PAH emission.

Correlation with charged PAHs. Bell et al. (2019) focussed on the 10-wide Galactic molecular ring surrounding the O-type star, λ-Orionis (cf. Fig. 3.45). We chose this region, because the Planck data indicate there is a large gradient of AME intensity. We fitted the spatially-resolved SED, at 0.25 angular sampling, using the AKARI 9 μm and IRAS 12 μm bands to constrain the PAH abundance, and longer wavelength bands for the rest of the emission. We used the dust SED code HerBIE (HiERarchical Bayesian Inference for dust Emission; Galliano, 2018, cf. Sect. 5.3.3). We were able to address the controversy, thanks to the combination of: (i)  rigorous SED fitting, allowing us to account for all the available information, not only a few broadbands; (ii) better observational constraint on the PAH emission (9 and 12 μm); and (iii) focussing on a clean region rather than the whole sky. Our results are shown in Fig. 3.46. We have found very good pixel-to-pixel correlations between the AME intensity, derived by Planck Collaboration et al. (2016a), and the dust and PAHs surface densities from SED fitting (e.g. cf. Fig. 3.46.a). Our analysis however show that, if the dust mass per pixel is very well correlated with the intensity of AME per U, IAMEU (ρ 0.88), the correlation is better with the mass of PAHs, and even better with the mass of ionized PAHs (ρ 0.92; cf. Fig. 3.46.b). Our Bayesian results also show that there is a 0 probability that the total dust could correlate better with AME than with PAHs. Our impression is that the study of Hensley et al. (2016) may have too quickly interpreted a poor correlation as an absence of causality. The scatter in WISE12μm was likely not the result of an intrinsic scatter of the PAH/AME correlation, but rather a combination of observational artefacts: (i) noise; (ii) contribution of the continuum and ionic lines. Our SED model and our better MIR coverage allowed us to more accurately recover the true PAH column density.

 AME correlates better with charged PAHs.

PIC
Figure 3.45: AKARI 9 μm map of λ-Orionis. This is a mosaic of broadband photometric images of PAHs within this Galactic molecular ring. Credit: Bell et al. (2019).
PIC PIC
Figure 3.46: AME correlation with ionized PAHs in λ-Orionis. Panel (a) shows the pixel correlation between IAMEU and MPAH+, derived from the SED modeling of (Bell et al., 2019). Each point and its uncertainty ellipse represent the posterior PDF of a pixel. Panel (b) shows the posterior PDFs of the correlation coefficients between two sets of parameters (such as panel a). The correlation between IAMEU and MPAH+ is the best of our study, and it is significantly better than with Mdust. Licensed under CC BY-SA 4.0.

3.3 Dust in Relation with the Gaseous and Stellar Contents

We end this chapter with a short introduction to ISMology, since the knowledge of the ISM as a whole is crucial to understanding ISD. We discuss a few of our results in this domain and illustrate the way dust can be used to better understand the gas. The video lectures and accompanying slides of the 2021 ISM of galaxies” summer school (35 hours of lecture in total), that we have organized, can provide a good introduction to this subject. Otherwise, the books of Spitzer (1978), Tielens (2005) and Draine (2011) are references.

3.3.1 The Phases of the ISM

The ISM is intrinsically a multi-phase medium, with large contrast densities and differences in temperatures. The order of magnitude of its average density is ngas 1 cm3. We list the physical characteristics of the main phases in Table 3.6. We will discuss each phase individually in the rest of this section.





Density

Temperature

Volume filling factor





HIM

3 × 103 cm3

106 K

50%





H II regions

1 105 cm3

104 K

1%





WIM

0.1 cm3

104 K

25%





WNM

0.3 cm3

104 K

30%





CNM

30 cm3

100 K

1%





Diffuse H2

100 cm3

50 K

0.1%





Dense H2

103 106 cm3

10 K

0.01%





Table 3.6: Phases of the ISM. Adapted from Tielens (2005) and Draine (2011). The sum of the filling factors is slightly larger than 100%, because these estimates are rough.

The cooling function. The way the gas cools across phases has a decisive impact on the multiphase structure of the ISM. It is possible to estimate its cooling rate as a function of temperature, or, in other words, how the thermal energy of the gas is converted into radiation. This quantity is called the cooling function. It is represented on Fig. 3.47. We have highlighted the dominant processes in the different temperature regimes.

The exact shape of this cooling function can vary sensibly. It depends on: (i) metallicity, as this parameter impacts the relative abundances of the different species; and (ii) the radiation field, which impacts the ionization state of the gas.

PIC
Figure 3.47: Cooling function of the ISM at Solar metallicity. The data below Tgas = 104 K is from Dalgarno & McCray (1972), and above, from Schure et al. (2009). We have highlighted the main cooling elements. In the neutral regime, we have assumed an electron fraction of x nenH = 104. Licensed under CC BY-SA 4.0.

3.3.1.1 The Neutral Atomic Gas

The neutral atomic gas is the most abundant phase in the MW: it accounts for 60% of the total ISM mass, and 8% of the total baryonic mass (stars and ISM). It fills up about 30% of the MW volume. The neutral gas is far from thermal equilibrium, but it is roughly in pressure equilibrium, with:

Pgas k = ngas × Tgas 3000K/cm3. (3.46)

The photoelectric heating. In neutral regions, the direct heating of the gas by absorption of stellar UV photons is not efficient, because only a small fraction of these photons can be absorbed through the different available atomic lines. In these regions, the heating of the gas is indirect. Dust grains absorb much more efficiently UV photons, with their spectrally continuous cross-section. The absorption of an energetic photon (of a few eV) can lead to the ejection of an electron, via the photoelectric effect. This electron will then collide with the gas and heat it. This is the photoelectric heating of the gas (de Jong, 1977; Draine, 1978; Bakes & Tielens, 1994; Weingartner & Draine, 2001b; Kimura, 2016). We have schematically represented this process in Fig. 3.48. The overall efficiency of this process is related to the integrated surface of dust grains. According to Table 2.3, this surface is dominated by small grains. The smallest grains, especially PAHs, are therefore responsible for most of this heating. Wolfire et al. (1995) give an expression for the photoelectric heating rate:

ngasΓ 𝜖PE(γ,Tgas) × G0 × ngas 1cm3 × 1025W/m3, (3.47)

where 𝜖PE is the efficiency of conversion of FUV energy into gas heating (it is a few percents), and G0 and γ have been defined in Eqs. (3.43-3.44).

PIC

Figure 3.48: Photoelectric heating in PDRs. We have implicitly assumed that the stellar power, L, is also the total power absorbed by dust, as PDRs are mostly optically thick at UV wavelengths. Licensed under CC BY-SA 4.0.

The two stable neutral atomic phases. For simplicity, let’s assume the heating of the neutral ISM is assured only by photoelectric heating. Let’s also use Eq. (3.47) with a fixed value of the photoelectric heating efficiency, 𝜖PE = 4.9%, and G0 = 1.7. The equilibrium is obtained when cooling and heating are balanced:

ngasΓ = ngas2Λ(T), (3.48)

which becomes, using Eq. (3.46):

Pgask = Γ × T Λ(T) . (3.49)

This is the black line we have represented in Fig. 3.49. There are three pressure equilibrium positions (the three dots), at the value of Eq. (3.46). We have hatched in grey the regime corresponding to unstable solutions. In this regime, the pressure indeed decreases with density. Thus, a small pressure increase, above the green dot, will decrease the density. It will thus make the gas expand and its temperature increase, moving further away from the green dot. On the opposite, a small decrease of the pressure, below the green dot, will increase the gas density, and make its temperature decrease, at the same time. The gas will thus collapse and move further away from the green dot. The two only stable solutions correspond to the two main neutral atomic phases of the ISM.

The Warm Neutral Medium
(WNM) is a diffuse phase with density of the order of ngas 0.3cm3 and temperature Tgas 104 K. It is heated essentially by the photoelectric effect, and also by cosmic rays. It cools via UV-optical emission lines, essentially Lyα121.6nm.
The Cold Neutral Medium
(CNM) is a denser phase with density of the order of ngas 30cm3 and temperature Tgas 100 K. It is heated essentially by the photoelectric effect, and also by cosmic rays. It cools essentially via fine-structure lines, [C II]158μm and also [O I]63μm.

Both of these phases are observed with H Iline, and UV-optical absorption lines.

 There are two distinct neutral atomic phases in pressure equilibrium: the WNM and the CNM.

PIC
Figure 3.49: Neutral ISM phase diagram. The black line is Eq. (3.49), with the cooling function of Fig. 3.47. The horizontal magenta line is the typical pressure of the ISM (Eq. 3.46). There are three pressure equilibrium solutions (the three dots), but one (the green dot) lies in the unstable regime (grey hatched). The arrows along each curve, near the dots, show the direction the gas would move if there was a perturbation around the equilibrium. Licensed under CC BY-SA 4.0.

3.3.1.2 The Ionized Gas

The ionized gas accounts for a moderate mass fraction of the ISM in the MW, but fills up a large volume. It is 23% of the ISM mass and 3% of the total baryonic mass (stars and ISM).

The Hot Ionized Medium (HIM). The HIM is a coronal gas (ngas 3 × 103cm3, Tgas 106 K; Table 3.6) filling up 50% of the MW volume. It is at pressure equilibrium with the WNM and CNM. McKee & Ostriker (1977) showed that this phase was heated by SN shockwaves. It is composed of overlapping SN bubbles, and is sometimes referred to as the intercloud medium. This gas, which pervades everywhere, can be found significantly above the Galactic disk, contrary to the other phase, which are contained to the disk. It is observed through FUV absorption lines of ions, and diffuse soft X-ray and synchrotron emissions.

H II regions. H II regions are short-lived dense ionized regions around OB star associations. The gas is ionized by photons from the massive stars. These regions are not in equilibrium, they are expanding. Strömgren (1939) has estimated the radius of an homogeneous sphere of ionized gas, around a star of H-ionizing photon rate, Q0. Such a sphere is called a Strömgren sphere. Its radius, Rs, can be estimated by balancing photoionization and recombination:

Q0 total ionization rate = 4π 3 Rs3n p numberofprotons × neαB recombination rate, (3.50)

where αB is the case B recombination rate. The product neαB is the electronic recombination rate to any H level n 2:

αB(Te) 2.6 × 1013T e34cm3s. (3.51)

Recombination down to n = 1 will indeed produce an ionizing photon that will be absorbed by another H atom. Rearranging Eq. (3.50), Strömgren’s radius is thus:

Rs = 3Q0 4πngas2αB 13. (3.52)

For an O5 star, with ngas = 103cm3, Rs 1 pc. This estimate can be refined, accounting for He and other atomic species, as well as dust screening (e.g. Osterbrock, 1989, for an extensive lecture).

The Warm Ionized Medium (WIM). The WIM is a diffuse phase (ngas 0.1cm3, Tgas 104 K; Table 3.6) filling up about 25% of the MW volume. It is roughly in pressure equilibrium with the other thermal phases, although it can be found expanding in some regions. This gas is photoionized by photons from OB star associations, escaping from H II regions. The electrons ejected by this photoionization provide the main heating source. It is observed through optical lines, essentially Hα656.3nm, as well as some fine-structure lines and free-free emission.

3.3.1.3 The Molecular Gas

The molecular gas constitutes only a moderate fraction of the ISM mass and a small fraction of its volume. It is however crucial, as it is where stars are formed. In the MW, 17% of the ISM mass is molecular, which corresponds to 2% of the total baryonic mass (stars and ISM).

Molecular hydrogen formation. The formation of the most abundant molecule of the Universe, H2, is inefficient in the gas phase. This is because the freshly formed molecule needs to evacuate 4.5 eV to remain stable. Yet, due to its symmetry, H2 does not have rotational transitions that would allow it to radiate at these energy levels. H2 formation thus takes place on grain surfaces (e.g. Bron et al., 2014, for a review). We have represented the two main processes on Fig. 3.50: the so-called Langmuir-Hinshelwood (physisorption) and Eley-Rideal (chemisorption) mechanisms. In both cases, the grain serves as a catalyst and helps evacuate the formation energy when the newly formed molecule is released in the gas phase. Similarly to the photoelectric heating, this process happening on grain surfaces, it takes place primarily on small grains. Temperature fluctuations therefore play an important role in its efficiency (e.g. Le Bourlot et al., 2012).

PIC

Figure 3.50: H2 formation on grain surface. Panel (a) illustrates the Langmuir-Hinshelwood mechanism, which is driven by physisorption. H atoms are first captured by the grains and migrate to form H2. The energy released by the formation liberates the molecule. Panel (b) demonstrates the Eley-Rideal mechanism, which is driven by chemisorption. In this case, an H atom is absorbed at a site where another H atom is already present. See Bron (2014) for more details. Licensed under CC BY-SA 4.0.

The diffuse molecular gas. Molecular gas can be observed at moderate densities (ngas 100cm3; Table 3.6). This is often associated with the CNM with large enough column densities to allow H2 to be self-shielded (i.e. its UV electronic lines are optically thick). It is also heated by photoelectric effect and cosmic rays. It cools primarily via [C II]158μm and can be observed through UV absorption lines.

Photodissociation regions. Same as H II regions, PDRs are a phase defined by the presence of massive stars in their vicinity. They are not a stable phase of the ISM, but they are very important since they are the interface between the H II region and the molecular cloud (e.g. Hollenbach & Tielens, 1997, for a review). Because of their high density and their high G0, they absorb a significant fraction of the FUV radiation by OB stars and reradiate it in the IR, as dust thermal emission and fine structure lines. Since they are the regions where molecules are being dissociated by FUV photons and subsequently recombine, they are the place of many important chemical reactions. Fig. 3.51 shows the structure of a typical PDR. We have performed an isobaric run with the Meudon PDR code (Le Petit et al., 2006), for Pgas = 107K/cm3 and G0 = 105. We show the variation of the abundances of the main species, and we represent in the upper part the H I/H2 and C II/C I/CO transitions. This figure demonstrates that H2, being efficiently self-shielded, exists at lower A(V), whereas CO appears deeper into the cloud.

PIC PIC

Figure 3.51: Structure of a PDR. The top drawing schematically represents the structure of a PDR. The ionizing stars are illuminating the cloud from the left. In the H II region, H is fully ionized. We enter the PDR at the ionization front. H is essentially neutral and atomic in this layer. We then pass the dissociation front around A(V ) 0.5 1, where H is essentially molecular (H2). Around A(V ) 1.5 2, C becomes neutral, and around A(V ) 3, CO becomes the dominant carbon-bearing species. We are at this stage in the cold molecular cloud. The two bottom panels show the results of an isobaric PDR model for Pgas = 107K/cm3 and G0 = 105 (using the Meudon PDR code;  Le Petit et al., 2006). Panel (a) shows the evolution as a function of A(V) of the densities of the most important species. These densities are normalized by the total H density, nH n(H+) + n(H0) + 2n(H 2). Panel (b) shows the evolution of the gas and dust temperatures. The exact densities, A(V) and temperatures are specific to the particular model we have run, but they give a rough idea of the typical values of these parameters. Licensed under CC BY-SA 4.0.

Dense molecular clouds. Dense molecular clouds (ngas 103 106cm3; Table 3.6) contain most of the molecular gas, concentrated in a small volume. These molecular clouds are gravitationally bound. Their structure is clumpy and filamentary. The gas motions are controlled by turbulence. They can be arranged in molecular complexes of sizes up to 100 pc (e.g. Solomon et al., 1987). The densest cores are collapsing and will lead to star formation. One of the most challenging issue of ISMology is the difficulty to measure the mass of molecular clouds. As we have mentioned earlier, H2 is a symmetric molecule. It thus does not have any dipole moment allowing rotational transitions. Its first transitions are its rovibrational levels (H extsubscript2,0-0S(0)28.3μm, H extsubscript2,0-0S(1)17.0μm and so on) that need temperatures of a few hundred K to be pumped. Cold molecular clouds are thus primarily traced by the second most abundant molecule, CO, which is asymmetric. CO rotational lines, 12CO(J =1 0) 2.6mm and 12CO(J =2 1) 1.3mm, are the most commonly observed transitions. The conversion of an observed 12CO(J =1 0) 2.6mm line intensity, ICO, to a H2 column density requires the knowledge of a CO-to-H2 conversion factor, XCO, such that (e.g. Bolatto et al., 2013):

N(H2) XCO × ICO. (3.53)

The XCO factor has been calibrated on Galactic molecular clouds: XCO 2 × 1020cm2(K.km/s)1. This value is however metallicity dependent, as we will see in Sect. 3.3.2.3.

3.3.2 Dust as a Diagnostic Tool

We have already discussed the potential of dust as a tracer of the physical conditions in the ISM, especially in Sect. 3.2.1.3. We review here a few examples where dust tracers were used to refine the results of studies dedicated to star formation or gas physics.

3.3.2.1 Dust to Study Star Formation

Star formation rates. Dust-related SFR tracers rely on the fact that young stars are extremely luminous and are enshrouded with dust. If the clouds are optically thick and if their covering factor is unity, the OB star luminosity is: LOB LTIR. Contrary to a common misconception, this is independent of dust properties. Assuming a typical IMF, burst age and metallicity, LOB can be converted to: SFR 1010 × L TIRL (e.g. Kennicutt, 1998a). The contribution of old stars to LTIR is negligible for high enough SFRs. Alternatively, monochromatic SFR indicators have been proposed. Calzetti et al. (2007) and Li et al. (2010) found that the 24 and 70 μm monochromatic luminosities were good local SFR indicators (on spatial scales of 0.5 1 kpc): SFR 2611 × [Lν(24μm)(LHz)]0.885 and SFR 1547 × Lν(70μm)(LHz). We have discussed the use of aromatic features as SFR tracers in Sect. 3.2.1.3. Finally, several composite indicators have been calibrated (Hao et al., 2011). By combining FUV or Hα656.3nm measurements with the 24 μm or TIR indicators, they account for the fact that star-forming regions are not completely opaque.

Resolving star formation. Hony et al. (2015) performed a comparison of different SFR estimators with the actual stellar content of the star-forming region N66, in the SMC. In this study, we derived the stellar surface density, Σ, based on individual star counts from HST photometry. When compared to the dust mass surface density, Σdust, derived from SED modeling, we found a significant scatter, at 6 pc linear resolution. The SFRs derived from Σdust or Hα656.3nm underestimate the more reliable, Σ-derived SFR, by up to a factor of 10. This is likely due to ionizing photons escaping the region 11. Finally, converting our Σdust map to a gas mass surface density map, Σgas, we found that the pixels of our region are lying above the Schmidt-Kennicutt relation 12. This might be due to low density gas, inefficient at forming stars, that is included in global Schmidt-Kennicutt relations, but absent when zooming on star-forming regions.

3.3.2.2 Photodissociation Regions

PDR Properties. We have participated to numerous studies aiming at constraining the PDR properties in low-metallicity environments (Cormier et al., 2010, 2012; Lebouteiller et al., 2012; Cormier et al., 2015; Chevance et al., 2016; Lebouteiller et al., 2017; Wu et al., 2018a; Cormier et al., 2019; Lebouteiller et al., 2019). The common point of these studies is their use of numerous fine structure lines observed by Herschel and Spitzer, as well as the dust emission, to constrain the main parameters of a PDR model (G0, ngas, filling factor). The challenge lies in the multiple degeneracies, due to the fact that a given line can trace several phases. For instance, [C II]158μm comes from the ionized gas, the neutral gas and from an important fraction of molecular clouds (Fig. 3.51). Such a degeneracy can be solved by using additional lines to constrain the properties of these different phases. Dust tracers are also useful, either as a constraint or as a self-consistency test. As an example, Chevance et al. (2016) modeled the gas properties in 30 Doradus and derived the typical depth of PDRs, in terms of visual extinction magnitude, APDR(V ). Fig. 3.52 shows the comparison of APDR(V ) to the value of this parameter, derived from SED modeling, Adust(V ). We can see that both quantities are in good agreement, validating the PDR results.

PIC PIC
(a) Three color image of 30 Doradus (b) A(V ) estimates
Figure 3.52: Comparison of visual extinctions in 30 Doradus. Panel (a) shows a map of 30 Doradus, seen through: (i)  [S IV]10.51μm (red; heavily-ionized gas;  Indebetouw et al., 2009); (ii) [Ne II]12.81μm (green; moderately ionized gas; Indebetouw et al., 2009); (iii) [C II]158μm (blue; neutral gas;  Chevance et al., 2016); and (iv) 12CO(J =3 2) 867μm (white contours; molecular gas; Minamidani et al., 2011). At the center of the image lies the SSC, R136. Panel (b) shows the comparison between the estimates of the visual extinction magnitudes, A(V ) (Eq. 2.4), in different regions of panel (a). The x-axis shows the A(V ) value derived from the PDR modeling of the available gas lines, whereas the y-axis shows the value of A(V ) inferred from dust SED modeling. Credit: (a) Chevance et al. (2016); (b) licensed under CC BY-SA 4.0.

Photoelectric heating. Assuming that [C II]158μm is the main gas coolant, the photoelectric efficiency, that we already discussed in Sect. 3.3.1.1, 𝜖PE, can be approximated by the gas-to-dust cooling ratio: 𝜖PE L[CII]LTIR. Detailed studies usually add other lines to the gas cooling rate, like [O I]63μm, to have a more complete estimate (e.g. Lebouteiller et al., 2012; Cormier et al., 2015; Lebouteiller et al., 2019). Overall, Smith et al. (2017) found that 0.1% 𝜖PE 1%, with an average of 𝜖PE 0.5%, in a sample of 54 nearby galaxies. It appears that 𝜖PE is lower when the dust temperature is higher (Rubin et al., 2009; Croxall et al., 2012). This is not likely the result of the destruction of the UIB carriers, as their intensity correlates the best with the [C II]158μm emission (e.g. Helou et al., 2001). It is rather conjectured to be due to the saturation of grain charging in UV-bright regions. The shape of the ISRF also has a consequence on the accuracy with which LTIR represents the true UV, photoelectric-efficient flux. Indeed, Kapala et al. (2017) showed that the variations of 𝜖PE in M 31 could be explained by the contribution of old stars to LTIR. Finally, one of the most puzzling features is that 𝜖PE is higher at low metallicity (Poglitsch et al., 1995; Madden et al., 1997; Cormier et al., 2015; Smith et al., 2017; Madden et al., 2020), while the UIB strength drops in these systems (Sect. 4.2.2.1). This is currently poorly understood. However, in the extreme case of I Zw 18 (Z 135Z), where no UIB is detected (e.g. Wu et al., 2006) and the photoelectric heating is estimated to be negligible, the gas-cooling-to-TIR ratio is still 1% (Lebouteiller et al., 2017). In this instance, we have shown the gas could be heated by X-rays.

3.3.2.3 The Molecular Gas and its Dark Layer

The dark gas. We have discussed in Sect. 3.3.1.3 that the photodissociation of H2 and CO at different depths into molecular clouds leads to biases in gas mass estimates. H2 is indeed self-shielded. It therefore exists at column densities roughly independent of metallicity. On the contrary, CO, which is significantly less abundant, is not self-shielded (i.e. its electronic lines remain optically thin at large column densities). Consequently, CO needs to be shielded by dust to survive. It thus exists only deeper into molecular clouds. The H2 gas that exists in regions where CO is photodissociated is often referred to as the CO-dark molecular gas. Other tracers can be used to constrain this dark gas: (i) dust emission (e.g. Israel, 1997; Leroy et al., 2011); (ii)  [C II]158μm (e.g. Madden et al., 1997); (iii) and γ-rays (e.g. Grenier et al., 2005). Using ICO to derive N(H2) with Eq. (3.53) can therefore be biased if the dark gas fraction is significantly larger than in the MW, where XCO has been calibrated. This is what happens in low-metallicity systems, where the dustiness is lower, because of the lower abundance of heavy elements (cf. Sect. 4.3.1). This is represented in Fig. 3.53. We see that in the low-metallicity cloud (on the right), CO cores are much smaller, because of the increased photodissociation of this molecule, compared to the left cloud (Solar metallicity). It results that the mass of CO-dark H2 is significantly larger at low metallicity. We have investigated this ISM component in the LMC, using dust mass surface density, concluding it could account between 10% and 100% of the total molecular gas mass (Galliano et al., 2011). Recently, we studied dark gas in the star-forming region N11 of the LMC, modeling the full set of IR emission lines (Lebouteiller et al., 2019). We showed that most of the molecular gas in this region is CO-dark and that [C II]158μm traces mostly this component. We extended this analysis to a sample of nearby dwarf galaxies (Madden et al., 2020). We found that 70 100% of the molecular gas mass is not traced by 12CO(J =1 0) 2.6mm.

 The molecular gas content of low-metallicity systems is dominated by dark gas.

PIC
Figure 3.53: Metallicity effect on the CO-dark gas. We have represented two molecular clouds photodissociated by nearby OB star associations. The left cloud has a Solar metallicity. Its CO core is efficiently shielded by dust. This is not the case for the right cloud, which has a low metallicity. In this cloud, the lower grain abundance causes a lower dust screening. Consequently, CO is photodissociated deeper into the cloud, whereas H2 remains self-shielded. Credit: Madden et al. (2020).

Pressure and radiation field. The pressure in molecular clouds can be significantly larger than in the pressure equilibrium phases of the ISM: Pgas 6 × 107K/cm3 in the Orion bar (Goicoechea et al., 2016); compared to Pgas 3000K/cm3 in the HIM, WIM and CNM (Eq. 3.46). We have studied the physical conditions of the molecular gas in the central region of the starbursting galaxy, M 83 (Wu et al., 2015). We used the CO Spectral Line Energy Distribution (SLED) observed by Herschel to estimate its column density and pressure, N(CO) and PCO, in different regions. We have also performed SED modeling to estimate the mean starlight intensity heating the grains, U (Eq. 3.38). This allowed us to show that both quantities are correlated (cf. Fig. 3.54.a). We also noted that the pressure gradient was oriented along a chain of radio sources, corresponding to a radio jet (Fig. 3.54.b). We derived a similar correlation between the ISRF strength and the gas thermal pressure in the Carina nebula (Wu et al., 2018a). Such a correlation was also found by Joblin et al. (2018) in the Orion bar. They argue that the photoevaporation of the PDR can explain this relation.

PIC PIC
(a) Pressure/ISRF relation (b) SFR map of M 83
Figure 3.54: Molecular gas pressure in the center of M 83. Panel (a) displays the relation between the average starlight intensity derived from SED modeling, and the molecular gas pressure, inferred from CO SLED modeling, in different regions in the center of M 83 (Wu et al., 2015). Panel (b) shows the SFR map of the central region of M 83. The color contours represent the molecular gas pressure, from PCO = 4 × 105K/cm3 (blue) to PCO = 2.4 × 106K/cm3 (red). The four diamonds indicates the position of the four radio sources reported by Maddox et al. (2006). Credit: (a) licensed under CC BY-SA 4.0; (b) Wu et al. (2015).

1.Stating that i=1NaiN i=1Nai1N, provided that ai 0, i.

2.VD99 solved this issue by assuming a power-law distribution of temperatures that we have discussed in Sect. 3.1.1.3. This is however an ad hoc solution that needs to be calibrated for each dust species, by using an actual MCRT model. Our goal being to demonstrate what a MCRT model brings, we chose to compare it only to the SED that the mega-grains approximation allows us to derive, by itself.

3.Literally for all wavelengths, from the X-rays to the radio.

4.Attenuation is not equivalent to extinction. The extinction is the sum of scattering and absorption along the sightline toward a point source, whereas the attenuation is a global account of the reddening of an ensemble of stars, potentially mixed with the dust. The extinction is directly linked to dust properties and is independent of geometry. The attenuation is a synthetic quantity depending both on the dust properties and on the ISM topology (e.g. Buat et al., 2019, for a recent discussion).

5.The Holmberg radius is the radius within which the B-band surface brightness of the galaxy is 26.5 magnitudes per squared arcsecond.

6.This factor is not precise, as the slope of the opacity is also changed. The difference in emissivity is thus not a sole scaling.

7.We note that the strength of the bump is controlled by the carbon-to-silicate grain ratio, a parameter that is poorly constrained by the SED fit.

8.The results of Madden et al. (2006) consisted in a first spectral decomposition of ISO/CAM spectra, that we have refined in Galliano et al. (2008b), and added Spitzer/IRS spectra to the sample.

9.γ-Fe2O3 is the notation to design ferromagnetic Fe2O3, called maghemite. It must be distinguished from its non-magnetic form, noted α-Fe2O3, called hematite. Both have distinct crystalline structures: cubic lattice for maghemite; trigonal crystals for hematite (e.g. Lȩcznar, 1977).

10.The radiance is the spectral integral of the specific intensity: RIνdν.

11.We saw a similar discrepancy in the center of M 83, where the [N II]122μm line was significantly more extended than other SFR tracers (Wu et al., 2015).

12.The Schmitt-Kennicutt relation is the empirical correlation between ΣSFR and Σgas for a wide diversity of galaxies (Schmidt, 1959; Kennicutt, 1998b).