ON RADIATION PRESSURE IN STATIC, DUSTY H ii REGIONS

Published 2011 April 25 © 2011. The American Astronomical Society. All rights reserved.
, , Citation B. T. Draine 2011 ApJ 732 100 DOI 10.1088/0004-637X/732/2/100

0004-637X/732/2/100

ABSTRACT

Radiation pressure acting on gas and dust causes H ii regions to have central densities that are lower than the density near the ionized boundary. H ii regions in static equilibrium comprise a family of similarity solutions with three parameters: β, γ, and the product Q0nrms; β characterizes the stellar spectrum, γ characterizes the dust/gas ratio, Q0 is the stellar ionizing output (photons/s), and nrms is the rms density within the ionized region. Adopting standard values for β and γ, varying Q0nrms generates a one-parameter family of density profiles, ranging from nearly uniform density (small Q0nrms) to shell-like (large Q0nrms). When Q0nrms ≳ 1052 cm−3 s−1, dusty H ii regions have conspicuous central cavities, even if no stellar wind is present. For given β, γ, and Q0nrms, a fourth quantity, which can be Q0, determines the overall size and density of the H ii region. Examples of density and emissivity profiles are given. We show how quantities of interest—such as the peak-to-central emission measure ratio, the rms-to-mean density ratio, the edge-to-rms density ratio, and the fraction of the ionizing photons absorbed by the gas—depend on β, γ, and Q0nrms. For dusty H ii regions, compression of the gas and dust into an ionized shell results in a substantial increase in the fraction of the stellar photons that actually ionize H (relative to a uniform-density H ii region with the same dust/gas ratio and density n = nrms). We discuss the extent to which radial drift of dust grains in H ii regions can alter the dust-to-gas ratio. The applicability of these solutions to real H ii regions is discussed.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Strömgren (1939) idealized photoionized nebulae around hot stars as static, spherical regions with a uniform density of ionized gas out to a bounding radius R. The Strömgren sphere model continues to serve as the starting point for studies of H ii regions around hot stars. However, a number of physical effects lead to departures from the simple Strömgren sphere model: dynamical expansion of the H ii region if the pressure in the surrounding neutral medium cannot confine the ionized gas, deviations from sphericity due to nonuniform density, motion of the star relative to the gas, injection of energy and momentum by a stellar wind, absorption of H-ionizing photons by dust grains, and radiation pressure acting on gas and dust. Each of these effects has been the object of a number of investigations, beginning with the study of ionization fronts by Kahn (1954).

Savedoff & Greene (1955) appear to have been the first to discuss the expansion of a spherical H ii region in an initially uniform neutral medium. Mathews (1967, 1969) and Gail & Sedlmayr (1979) calculated the dynamical expansion of an H ii region produced by an O star in a medium that was initially neutral, including the effects of radiation pressure acting on the dust. Mathews (1967, 1969) showed that radiation pressure on dust would produce low-density central cavities in H ii regions. More recently, Krumholz & Matzner (2009) re-examined the role of radiation pressure on the expansion dynamics of H ii regions, concluding that radiation pressure is generally unimportant for H ii regions ionized by a small number of stars, but is important for the expansion dynamics of giant H ii regions surrounding clusters containing many O-type stars. Their study concentrated on the forces acting on the dense shell of neutral gas and dust bounding the H ii region; hence they did not consider the density structure within the ionized region.

Dust absorbs hν>13.6 eV photons that would otherwise be able to ionize hydrogen, thereby reducing the extent of the ionized zone. Petrosian et al. (1972) developed analytic approximations for dusty H ii regions. They assumed the gas density to be uniform, with a constant dust-to-gas ratio, and found that dust could absorb a substantial fraction of the ionizing photons in dense H ii regions. Petrosian et al. did not consider the effects of radiation pressure.

Dopita et al. (2003, 2006) constructed models of compact H ii regions, including the effects of radiation pressure on dust, and presented models for different ionizing stars and bounding pressures. In these models, radiation pressure produces a density gradient within the ionized gas.

This paper provides a systematic discussion of the structure of dusty H ii regions that are assumed to be in equilibrium with an external bounding pressure. The assumptions and governing equations are presented in Section 2, where it is shown that dusty H ii regions are essentially described by a three-parameter family of similarity solutions. In Section 3, we show density profiles for selected cases as well as surface brightness profiles. The characteristic ionization parameter U1/2 and the fraction (1 − fion) of the ionizing photons absorbed by dust are calculated. Dust grain drift is examined in Section 4, where it is shown that it can alter the dust-to-gas ratio in the centers of high-density H ii regions. The results are discussed in Section 5 and summarized in Section 6.

2. EQUILIBRIUM MODEL

Consider the idealized problem of a static, spherically symmetric equilibrium H ii region ionized by a point source, representing either a single star or a compact stellar cluster. Assume a constant dust-to-gas ratio (the validity of this assumption will be examined later). For simplicity, ignore scattering, and assume σd, the dust absorption cross section per H nucleon, to be independent of photon energy hν over the ∼5–30  eV range containing most of the stellar power.

Let the star have luminosity L = Ln + Li = L39 ×  1039 erg s−1, where Ln and Li are the luminosities in hν < 13.6 eV and hν>13.6 eV photons, respectively. The rate of emission of hν>13.6 eV photons is Q0 ≡ 1049Q0,49 s−1 and the mean energy of the ionizing photons is 〈hν〉iLi/Q0. A single main-sequence star of spectral type O6V has L39 = 0.80 and Q0,49 = 0.98 (Martins et al. 2005). A compact cluster of OB stars might be treated as a point source with much larger values of Q0,49 and L39.

Ignore He, and assume the H to be nearly fully ionized, with photoionization balancing "Case B" radiative recombination, with "on-the-spot" absorption of hν>13.6 eV recombination radiation. Take the effective radiative recombination coefficient to be αB ≈ 2.56 × 10−13T−0.834 cm3 s−1 for 0.5 ≲ T4 ≲ 2, with T4T/104 K, where T is the gas temperature.

Assume the gas to be in dynamical equilibrium (the neutral gas outside the ionized zone is assumed to provide a confining pressure). Static equilibrium then requires that the force per unit volume from radiation pressure be balanced by the pressure gradient:

Equation (1)

where n(r) is the proton density, Liϕ(r) is the power in hν>13.6 eV photons crossing a sphere of radius r, and τ(r) is the dust absorption optical depth. Equation (1) underestimates the radiation pressure force, because it assumes that recombination radiation (including Lyα) and cooling radiation escape freely.

The functions ϕ(r) and τ(r) are determined by

Equation (2)

Equation (3)

with boundary conditions ϕ(0) = 1 and τ(0) = 0. Define a characteristic density and length scale

Equation (4)

Equation (5)

and the dimensionless parameters

Equation (6)

Equation (7)

The parameter β, the ratio of the power in non-ionizing photons to the power in photons with hν>13.6 eV, depends solely on the stellar spectrum. We take β = 3 as our standard value, corresponding to the spectrum of a T = 32,000 K blackbody, but we also consider β = 2 (T = 45,000 K) and β = 5 (T = 28,000 K); the latter value might apply to a cluster of O and B stars.

Momentum can be transferred to a dust grain by photon absorption, but also by scattering. The cross section σd appearing in Equation (1) should be 〈σpr〉, the radiation pressure cross section per H, σpr(ν) ≡ σabs + (1 − 〈cos θ〉)σsca, averaged over the spectrum of the radiation field at radius r, where σabs(ν) and σsca(ν) are the absorption and scattering cross section per H, respectively, and 〈cos θ〉 is the mean value of the cosine of the scattering angle θ for photons of frequency ν.

In Equations (2) and (3), σd characterizes the effectiveness of the dust in attenuating the radiation field. While scattering does not destroy the photon, it does increase the probability of the photon undergoing subsequent absorption. Thus, the value of σd in Equations (2) and (3) should exceed 〈σabs〉.

Figure 1(a) shows the dust absorption cross section per H nucleon averaged over a blackbody spectrum, for two dust models (Weingartner & Draine 2001; Zubko et al. 2004) that reproduce the wavelength-dependent extinction in the diffuse interstellar medium (ISM) using mixtures of polycyclic aromatic hydrocarbons (PAHs), graphite, and amorphous silicate grains. Figure 1(b) shows that 〈σpr〉, the radiation pressure cross section averaged over blackbody spectra, is only slightly larger than 〈σabs〉. Given the uncertainties in the nature of the dust in H ii regions, it is reasonable to ignore the distinction between 〈σpr〉 and the attenuation cross section and simply take σd = 〈σpr〉 in Equations (1)–(3).

Figure 1.

Figure 1. (a) Absorption cross section per H and (b) radiation pressure cross section per H, averaged over blackbody spectra, as functions of the blackbody temperature T, for the dust models of Weingartner & Draine (2001, WD01) and Zubko et al. (2004, ZDA04). Broken lines show averages over hν < 13.6 eV only, appropriate for dust in neutral gas.

Standard image High-resolution image

For dust characteristic of the diffuse ISM, one could take $\langle \sigma _{\rm pr}\rangle \approx 1.5\times 10^{-21}\,{\rm cm}^2{\,\rm H}^{-1}$ for 2.5 × 104 K ≲ Trad ≲ 5 × 104 K. However, dust within an H ii region may differ from average interstellar dust. For example, the small-size end of the size distribution might be suppressed, in which case σd would be reduced. Low-metallicity galaxies will also have lower values of σd, simply because there is less material out of which to form grains. In this work, we will assume a factor ∼1.5 reduction in σd relative to the local diffuse ISM, taking $\sigma _d\approx 1\times 10^{-21}\,{\rm cm}^2{\,\rm H}^{-1}$ as the nominal value, but larger and smaller values of σd will also be considered.

The dimensionless parameter γ (defined in Equation (7)) depends also on the gas temperature T and on the mean ionizing photon energy $\langle h\nu \rangle {\rm_i}$, but these are not likely to vary much for H ii regions around OB stars. We take γ = 10 as a standard value, but will also consider γ = 5 and γ = 20. Low-metallicity systems would be characterized by small values of γ.

Switching to dimensionless variables yr0 and un0/n, the governing Equations (1)–(3) become

Equation (8)

Equation (9)

Equation (10)

with initial conditions ϕ(0) = 1 and τ(0) = 0. The solutions are defined for 0 < yymax, where ymax is determined by the boundary condition ϕ(ymax) = 0. The actual radius of the ionized zone is R = ymaxλ0. For each solution u(y), the mean density is

Equation (11)

the rms density is

Equation (12)

and the gas pressure at the edge of the H ii region is

Equation (13)

Let

Equation (14)

be the radius of a dustless Strömgren sphere with density nrms = 103nrms,3 cm−3. The fraction of the hν>13.6 eV photons that are absorbed by H is simply

Equation (15)

For given (β, γ), varying the initial value1 of u = n0/n at some fixed y = r0 generates solutions with different density profiles. Therefore, the full set of solutions forms a three-parameter family of similarity solutions u(y), ϕ(y), and τ(y), parameterized by β, γ, and a third parameter. The third parameter can be taken to be Q0nrms. For dusty H ii regions, an alternative choice for the third parameter is the dust optical depth on a path Rs0 with density nrms:

Equation (16)

Equation (17)

The static H ii regions described by Equations (1)–(3) are determined by seven distinct dimensional quantities: three parameters describing the central star (Q0, $\langle h\nu \rangle {\rm_i}$, and Ln), the recombination rate coefficient αB, the thermal energy kT, the dust cross section per nucleon σd, and the external pressure pedge confining the H ii region. According to the present analysis, this seven-parameter family of solutions actually reduces to a three-parameter family of similarity solutions. The dimensionless parameters β and γ, plus the choice of an initial value for the function u near y = 0, suffice to determine the scaled density profile n(r)/n0 and radius ymax = R0: this is the three-parameter family of similarity solutions.

Specifying numerical values for the ratios Q0B and $kT/(\alpha _B\langle h\nu \rangle {\rm_i})$ fixes the values of n0 and λ0, thus giving n(r) for r < R. Thus far we have invoked five independent parameters, but have not actually specified either kT or αB.

Specifying kT and αB—the sixth and seventh parameters— allows us to compute the actual values of Q0 and $\langle h\nu \rangle {\rm_i}$, and the bounding pressure pedge = 2n(R)kT.

If the "initial value" of u near the origin is taken as a boundary condition, then pedge emerges as a derived quantity. However, if we instead regard pedge as a boundary condition, then the initial value of u ceases to be a free parameter, and instead must be found (e.g., using a shooting technique) so as to give the correct boundary pressure pedge: the initial value of u is thus determined by the seven physical parameters (Q0, $\langle h\nu \rangle {\rm_i}$, Ln, αB, T, σd, and pedge).

Thus, we see how the three-parameter family of dimensionless similarity solutions corresponds to a seven-parameter family of physical solutions.

3. RESULTS

Equations (8)–(10) can be integrated numerically. Figure 2(a) shows the solution for the case where no dust is present (γ = 0). Radiation pressure associated with photoionization produces a density gradient in the H ii region, but it is modest unless Q0nrms is very large. The central density is nonzero. For Q0,49nrms ≲ 103 cm−3, the density is uniform to within ±15%.

Figure 2.

Figure 2. Normalized density profiles of static equilibrium H ii regions, as a function of r/R, where R is the radius of the ionized region. Profiles are shown for seven values of Q0nrms; the numerical values given in the legends assume T4 = 0.94 and $\langle h\nu \rangle {\rm_i}=18\,{\rm eV}$. (a) Dustless (γ = 0), (b) γ = 5, (c) γ = 10, and (d) γ = 20.

Standard image High-resolution image

As discussed above, the dust abundance relative to H is characterized by the parameter γ. Density profiles are shown in Figures 2(b)–(d) for β = 3 and γ = 5, 10,  and 20, corresponding approximately to σd = 0.5 × 10−21, 1 × 10−21,  and 2 × 10−21 cm2. With dust present, the density formally goes to zero at r = 0. For fixed γ, the size of the low-density central cavity (as a fraction of the radius R of the ionization front) increases with increasing Q0nrms. The enhancement of the density near the ionization front also becomes more pronounced as Q0nrms is increased. For β = 3, γ = 10, and Q0,49nrms = 105 cm−3, we find n(R) = 2.5nrms.

The state of ionization of the gas is determined by the hardness of the radiation field and the value of the dimensionless "ionization parameter"

Equation (18)

where n(hν>IH) is the density of photons with hν>IH. Within an H ii region, the value of U varies radially. As a representative value, we evaluate U1/2, the value at the "half-ionization" radius R1/2, the radius within which 50% of the H ionizations and recombinations take place.2 In a uniform-density dustless H ii region, R1/2 = 2−1/3Rs0 is the same as the half-mass radius, and

Equation (19)

For our present models,

Equation (20)

where y1/2 = R1/20 is the value of y within which 50% of the H ionizations and recombinations take place. Figure 3 shows U1/2 as a function of Q0nrms for static dusty H ii regions with radiation pressure, for selected values of β and γ. For small values of Q0nrms, dust and radiation pressure are negligible and U1/2 coincides with U(no dust)1/2 (Equation (19)). However, for large values of Q0nrms, U1/2 falls below U(no dust)1/2. For γ ≈ 10—corresponding to the dust abundances that we consider to be likely for Galactic H ii regions—we see that U1/2 ≈ 0.07 ± 0.02 for Q0,49nrms ≳ 104 cm−3.

Figure 3.

Figure 3. Ionization parameter U1/2 at the half-ionization radius in dusty H ii regions (see the text), calculated assuming T4 = 0.94 and $\langle h\nu \rangle {\rm_i}=18\,{\rm eV}$.

Standard image High-resolution image

The emission measure EM(b) = ∫n2eds is shown as a function of impact parameter b in Figure 4. For small values of Q0nrms, the intensity profile is close to the semicircular profile of a uniform-density sphere. As Q0nrms is increased, the profile becomes flattened, but, if no dust is present (γ = 0, Figure 2(a)), the ionized gas only begins to develop an appreciable central minimum for Q0,49nrms ≳ 104.5 cm−3.

Figure 4.

Figure 4. Normalized emission measure (EM) profiles for a cut across the center of H ii regions with (a) γ = 0 (no dust), (b) γ = 5, (c) γ = 10, and (d) γ = 20. Profiles are shown for selected values of Q0nrms. Numerical values of Q0,49nrms assume T4 = 0.94 and $\langle h\nu \rangle {\rm_i}=18\,{\rm eV}$.

Standard image High-resolution image

When dust is present, however, the profiles are strongly affected. For standard parameters β = 3 and γ = 10, the EM shows a pronounced central minimum for Q0,49nrms ≳ 103 cm−3, with a peak-to-minimum ratio >2 for Q0,49nrms ≳ 104 cm−3. As Q0nrms is increased, the ionized gas becomes concentrated in a thin, dense shell, the peak intensity near the edge rises, and the central EM changes from EM(0) = 2n2rmsR for small Q0nrms to EM(0) → (2/3)n2rmsR as Q0nrms.

When τd0 ≫ 1, the present models have the ionized gas concentrated in a thin, dense shell. Figure 5(a) shows the ratio of the rms density nrms to the mean density 〈n〉 as a function of τd0. The highest ionized density occurs at the outer edge of the ionized zone, and Figure 5(b) shows the ratio nedge/nrms as a function of τd0.

Figure 5.

Figure 5. For dusty H ii regions with γ = 5, 10, 20 and β = 2, 3, 5, as a function of τd0: (a) ratio nrms/〈n〉 of rms density to mean density; (b) ratio n(R)/nrms of the edge density to the rms density; (c) center-to-edge dust optical depth τ(R); (d) ratio of peak EM/central EM. Results are for $\langle h\nu \rangle {\rm_i}=18\,{\rm eV}$.

Standard image High-resolution image

In the low-density limit Q0nrms → 0, the dust optical depth from center to edge τ(R) ≈ τd0. The actual dust optical depth from center to edge is shown in Figure 5(c). When the H ii region develops a dense shell, which occurs for τd0 ≳ 3, the actual dust optical depth τ(R) is significantly smaller than the value τd0. Figure 5(c) shows that for τd0 = 40, for example, the actual dust optical depth τ(R) is only in the range 1–2.3, depending on the values of β and γ.

The shell-like structure is also apparent in the ratio of the peak intensity to the central intensity. As seen in Figures 4(b)–(d), dust causes the peak intensity to be off-center. For fixed β and γ, the ratio of peak intensity to central intensity, I(peak)/I(center), increases monotonically with increasing τd0, as shown in Figure 5(d).

Because the shell is dense, radiative recombination is rapid, the neutral hydrogen fraction is enhanced, and H atoms can compete with dust to absorb hν>13.6 eV photons. Figure 6 shows fion, the fraction of the hν>13.6 eV photons emitted by the star that photoionize H (i.e., are not absorbed by dust), as a function of the parameter τd0. Results are shown for β = 2, 3, 5 and γ = 5, 10, 20. For 2 ⩽ β ⩽ 5, 5 ⩽ γ ⩽ 20, and 0 ⩽ τd0 ⩽ 40, the numerical results in Figure 6 can be approximated by the fitting formula

Equation (21)

Equation (22)

Equation (23)

where β, γ, and τd0 are given by Equations (6), (7), and (16). The form of Equations (21)–(23) has no physical significance, but Equation (21) can be used to estimate the total H ionization rate fionQ0 in dusty H ii regions.

Figure 6.

Figure 6. Fraction fion of the hν>13.6 eV photons that photoionize H in dusty H ii regions with radiation pressure, as a function of τd0, for β = 2, 3, 5 and γ = 5, 10, 20. Calculated assuming $\langle h\nu \rangle {\rm_i}=18\,{\rm eV}$. The dotted lines show the fitting formula (21) for each case. Also shown is fion calculated for assumed uniform density (Petrosian et al. 1972).

Standard image High-resolution image

Even for large values of τd0, Figure 6 shows that ∼1/3 of the hν>13.6 eV photons are absorbed by hydrogen. This contrasts with the uniform-density models of Petrosian et al. (1972), where the fraction of the hν>13.6 eV photons that are absorbed by the gas goes to zero as τd0 becomes large.

4. DUST DRIFT

4.1. Gas Drag versus Radiation Pressure

Equation (1) assumes the dust to be tightly coupled to the gas, so that the radiation pressure force on the dust can be considered to act directly on the gas. In reality, radiation pressure will drive the dust grains through the plasma. If the grains approach their terminal velocities (i.e., acceleration can be neglected), then, as before, it can be assumed that the radiation pressure force is effectively applied to the gas. However, the motion of the dust grains will lead to changes in the dust/gas ratio, due to the movement of the grains from one zone to another, as well as because of grain destruction. Here, we estimate the drift velocities of grains.

Let Qprπa2 be the radiation pressure cross section for a grain of radius a. Figure 7 shows Qpr(a, λ) averaged over blackbody radiation fields with T = 25,000 K, 32,000 K, and 40,000 K, for carbonaceous grains and amorphous silicate grains. For spectra characteristic of O stars, 〈Qpr〉 ≈ 1.5 for 0.02 μm ≲ a ≲ 0.25 μm.

Figure 7.

Figure 7. Spectrum-averaged radiation pressure efficiency factor 〈Qpr〉 as a function of radius, for T = 25,000 K, 32,000 K, and 40,000 K blackbody spectra. For temperatures characteristic of O stars, 〈Qpr〉 ≈ 1.5 to within 20% for 0.02 μm ≲ a ≲ 0.25 μm.

Standard image High-resolution image

If the magnetic field B = 0, the terminal velocity vd for a grain at distance r is determined by balancing the forces due to radiation pressure and gas drag:

Equation (24)

where the drag function G(s), including both collisional drag and Coulomb drag, can be approximated by (Draine & Salpeter 1979)

Equation (25)

Equation (26)

where U is the grain potential, a−5a/10−5 cm, and n3ne/103 cm−3. The drag force from the electrons is smaller than that from the ions by at least $\sqrt{m_e/m_p}$, and can be neglected. The charge on the grains will be determined by collisional charging and photoelectric emission. Collisional charging would result in eU/kT ≈ −2.51 (Spitzer 1968), or U ≈ −2.16T4 V. Photoelectric charging will dominate close to the star, but is expected to result in potentials U ≲ 10 V. Taking |eU/kT| ≈ 2.5 and ln Λ ≈ 14.8 as representative,

Equation (27)

Note that G(s) is not monotonic: as s increases from 0, G(s) reaches a peak value ∼42 for s ≈ 0.89, but then begins to decline with increasing s as the Coulomb drag contribution falls. At sufficiently large s, the direct collisional term becomes large enough that G(s) rises above ∼42 and continues to rise thereafter.

The drag time for a grain of density ρ = 3 g cm−3 in H ii gas is

Equation (28)

For n3 ≳ 0.01, this is sufficiently short that each grain can be assumed to be moving at its terminal velocity vd, with the isothermal Mach number $s\equiv v_d/\sqrt{2kT/m_{\rm H}}$ determined by the dimensionless equation

Equation (29)

Equation (29) is solved to find s(r). For 20 < G < 42, there are three values of s for which the drag force balances the radiation pressure force. The intermediate solution is unstable; we choose the smaller solution,3 which means that s undergoes a discontinuous jump from ∼0.9 to 6.2 at G ≈ 42. The resulting terminal velocity v(r) is shown in Figure 8 for seven values of Q0,49nrms. The velocities in the interior can be very large, but the velocities where most of the dust is located (τ(r)/τ(R)>0.5) are much smaller.

Figure 8.

Figure 8. Radial drift velocities vd,r for six different H ii regions, all with β = 3 and γ = 10, for T4 = 0.94, $\langle h\nu \rangle {\rm_i}=18\,{\rm eV}$, Qpr = 1.5, |eU/kT| = 2.5, and B = 0. (a) vd,r vs. r/R. All solutions have large drift velocities near the center, which will result in removal of the dust from the interior. Drift velocities increase with increasing Q0nrms. (b) vd,r as a function of dust column density τ(r). Even if B = 0 or Br, drift velocities vd>75 km s−1 occur only in a region with a small fraction of the dust. In most of the volume, drift will not result in grain destruction.

Standard image High-resolution image

In the outer part of the bubble, where most of the gas and dust are located, the drift velocities are much more modest. This is visible in Figure 8(a), where the drift speeds become small as rR, but is more clearly seen in Figure 8(b), showing drift speeds as a function of normalized optical depth. The range 0.5 < τ(r)/τ(R) < 1 contains more than 50% of the dust, and throughout this zone the drift speeds are ≲0.3 km s−1 even for Q0,49nrms as large as 107 cm−3. With drift speeds vd ≲ 0.3 km s−1, grains will not be destroyed, except perhaps by shattering in occasional collisions between grains with different drift speeds. However, for large values of nrms, these grains are located close to the boundary, the drift times may be short, and the grains may be driven out of the H ii and into the surrounding shell of dense neutral gas. This will be discussed further below.

4.2. Magnetic Fields

Let epsilonBB2/16πnkT be the ratio of magnetic pressure to gas pressure. The importance of magnetic fields for the grain dynamics is determined by the dimensionless ratio ωτdrag, where ω ≡ QB/Mc is the gyrofrequency for a grain with charge Q and mass M in a magnetic field B:

Equation (30)

If |eU/kT| ≈ 2.5 and ln Λ ≈ 15, then (G(s)/s) ≈ 71 for s ≲ 0.5.

Let the local magnetic field be ${\bf B}=B(\hat{{\bf r}}\cos \theta +\hat{{\bf y}}\sin \theta)$. The steady-state drift velocity is

Equation (31)

Equation (32)

Equation (33)

Equation (34)

where vd,r, vd,y, and vd,z are the radial and transverse components. If sin θ → 0, the magnetic field does not affect the radiation-pressure-driven drift velocity, but in the limit sin θ → 1 magnetic effects can strongly suppress the radial drift if ωτdrag ≫ 1 and cos θ ≪ 1.

The magnetic field strength is uncertain, but it is unlikely that the magnetic energy density will be comparable to the gas pressure; hence, epsilonB ≲ 0.1. From Equation (30), it is then apparent that if the magnetic field is strong (epsilonB ≈ 0.1), magnetic effects on the grain dynamics can be important in low-density H ii regions, but will not be important for very high densities: (ωτdrag)2 ≲ 1 for n3 ≳ 170a−2−5epsilonB.

4.3. Drift Timescale

When radiation pressure effects are important, the gas and dust are concentrated in a shell that becomes increasingly thin as Q0nrms is increased. The drift velocities where most of the dust is located are not large (see Figure 8(b)), but the grains are also not far from the ionization front. The timescale on which dust drift would be important can be estimated by calculating the drift velocity at the radius r0.5 defined by τ(r0.5) = 0.5τ(R). More than 50% of the dust has r0.5 < r < R. Figure 9 shows the characteristic drift time

Equation (35)

If no magnetic field is present, the drift velocity depends only on T and the dimensionless quantities {ϕ, τ, u, y, 〈Qpr〉} (see Equation (29)). It is easy to see that for fixed T and Q0nrms, the radius RQ0; thus tdriftQ0. Figure 9 shows tdrift/Q0,49. For Q0,49 = 1, H ii regions with nrms>103 cm−3 have tdrift < 106 yr if magnetic effects are negligible. If a magnetic field is present with Br and epsilonB = 0.1, then the grain drift is slowed, but drift times of <1 Myr are found for nrms>104 cm−3. Therefore, compact and ultracompact H ii regions around single O stars are able to lower the dust/gas ratio by means of radial drift of the dust on timescales of ≲ Myr. However, if the O star is moving relative to the gas cloud with a velocity of more than a few  km s−1, then individual fluid elements pass through the ionized zone on timescales that may be shorter than the drift timescale, precluding substantial changes in the dust-to-gas ratio.

Figure 9.

Figure 9. Drift timescale tdrift/Q0,49 (see Equation (35)) for β = 3 and γ = 5, 10, and 20 (assuming T4 = 0.94, $\langle h\nu \rangle {\rm_i}=18 \,{\rm eV}$). The dust grains are assumed to have 〈Qpr〉 = 1.5 and |eU/kT| = 2.5. Solid lines are for B = 0 (or Br). Broken lines are for a = 0.03 μm, and epsilonBQ0,49 = 0.1 and 102.

Standard image High-resolution image

Grain removal by drift can also occur for giant H ii regions. As an example, consider a giant H ii region ionized by a compact cluster of ∼103 O stars emitting ionizing photons at a rate Q0 = 1052 s−1. For nrms = 103 cm−3, we have Q0,49nrms = 106 cm−3, and we see that if B = 0, the drift timescale is only tdrift ≈ 2 × 105 yr. If a magnetic field is present with Br and epsilonB = 0.1, then from Figure 9 the drift timescale tdrift is increased, but only to ∼106 yr. It therefore appears possible for radiation-pressure-driven drift to remove dust from giant H ii regions provided they are sufficiently dense.

Aside from magnetic effects, the drift speeds at a given location depend only on 〈Qpr〉 and T4 (see Equation (29)). Figure 7 shows that 〈Qpr〉 is constant to within a factor ∼1.5 for a ≳ 0.010 μm. Hence, radiation-pressure-driven drift would act to drive grains with a ≳ 0.01 μm outward. Smaller grains will drift as well, but more slowly. Because of this, the gas-to-dust ratio in the centers of H ii regions should in general be lower than the gas-to-dust ratio in the gas prior to ionization. The dust-to-gas ratio will first be reduced in the center, where the drift speeds (see Figure 8) are large. Dust drift will also alter the dust-to-gas ratio in the outer ionized material, initially raising it by moving dust outward from the center. In an initially uniform neutral cloud, the ionization front expands rapidly at early times (see, e.g., Figure 37.3 in Draine 2011) but in gas with n3 ≳ 1, at late times the ionization front will slow to velocities small enough for dust grains to actually drift outward across the ionization front, lowering the overall dust-to-gas ratio within the H ii region.

4.4. Grain Destruction

Arthur et al. (2004) computed models of uniform-density H ii regions including the effects of dust destruction by sublimation or evaporation, finding that the dust/gas ratio can be substantially reduced near the star. If the maximum temperature at which a grain can survive is Tsub, and the Planck-averaged absorption efficiencies are Quv and Qir for T = T and T = Tmax, then grains will be destroyed within a distance rsub with

Equation (36)

For parameters of interest (e.g., L39/Q0,49 ≈ 1, L39 ≲ 102) we find rsub/Rs0 ≪ 1 for nrms ≲ 105 cm−3, and sublimation would therefore destroy only a small fraction of the dust.

As we have seen, we expect radiation pressure to drive grains through the gas, with velocity given by Equation (31). Drift velocities vd ≳ 75 km s−1 will lead to sputtering by impacting He ions, with sputtering yield Y(He) ≈ 0.2 for 80 ≲ v ≲ 500 km s−1 (Draine 1995). For hypersonic motion, the grain of initial radius a will be destroyed after traversing a column density

Equation (37)

for a grain density ρ/μ = 1 × 1023 cm−3, appropriate for either silicates (e.g., FeMgSiO4, ρ/μ ≈ 3.8 g cm−3/25mH = 9 × 1022 cm−3) or carbonaceous material (2 g cm−3/12mH = 1.0 × 1023 cm−3). Therefore, the dust grain must traverse material with (initial) dust optical depth

Equation (38)

if it is to be substantially eroded by sputtering. However, Figure 8(b) shows that even in the absence of magnetic effects, vd ≳ 75 km s−1 occurs only in a central region with τd < 0.05. Therefore, sputtering arising from radiation-pressure-driven drift will not appreciably affect the dust content.

5. DISCUSSION

5.1. Absorption of Ionizing Photons by Dust

For a sample of 13 Galactic H ii regions, Inoue (2002) used infrared and radio continuum observations to obtain the values of fion shown in Figure 10. The estimated values of fion are much larger than would be expected for uniform H ii regions with dust-to-gas ratios comparable to the values found in neutral clouds. Inoue (2002) concluded that the central regions of these H ii regions must be dust-free, noting that this was likely to be due to the combined effects of stellar winds and radiation pressure on dust. As seen in Figure 10, the values of fion found by Inoue are entirely consistent with what is expected for static H ii regions with radiation pressure for 5 ≲ γ ≲ 20 (corresponding to 0.5 ≲ σd,−21 ≲ 2), with no need to appeal to stellar winds or grain destruction.

Figure 10.

Figure 10. Photoionizing fraction fion for 12 Galactic H ii regions, as estimated by Inoue (2002) from infrared and radio observations, vs. Q0,49neT0.834 (see the text). fion cannot exceed 1; therefore the high value found for G298.22−0.34 gives some indication of the uncertainties in estimation of fion. Solid lines: fion for H ii regions with radiation pressure for dust characterized by γ = 5, 10, and 20. Broken line: fion for uniform H ii regions with σd = 10−21 cm2 H−1.

Standard image High-resolution image

5.2. The Density–Size Correlation for H ii Regions

ii regions come in many sizes, ranging from H ii regions powered by a single O star to giant H ii regions ionized by a cluster of massive stars. The physical size of the H ii region is obviously determined both by the total ionizing output Q0 provided by the ionizing stars, and the rms density nrms of the ionized gas, which is regulated by the pressure pedge of the confining medium. With the balance between photoionization and recombination determining the size of an H ii region, an anticorrelation between size D and density nrms is expected, and was observed as soon as large samples of H ii regions became available (e.g., Habing & Israel 1979; Kennicutt 1984). For dustless H ii regions, one expects nrmsD−1.5 for fixed Q0, but for various samples relations close to nrmsD−1 were reported (e.g., Garay et al. 1993; Garay & Lizano 1999; Kim & Koo 2001; Martín-Hernández et al. 2005). For ultracompact H ii regions, Kim & Koo (2001) attribute the neD−1 trend to a "champagne flow" and the hierarchical structure of the dense gas in the star-forming region, but Arthur et al. (2004) and Dopita et al. (2006) argue that the neD−1 trend is a result of both absorption by dust and radiation pressure acting on dust in static H ii regions.

Hunt & Hirashita (2009) recently re-examined the size–density relationship. They interpreted the size–density relation for different observational samples in terms of models with different star formation rates (and hence different time evolution of the ionizing output Q0(t)), and differences in the density of the neutral cloud into which the H ii region expands. Their models did not include the effects of radiation pressure on dust; at any time the ionized gas in an H ii region was taken to have uniform density, resulting in overestimation of the dust absorption.

Figure 11(a) shows a grid of nrms versus D for the present models, for four combinations of (β, γ). While differences between the models with different (β, γ) can be seen, especially for high Q0 and high pedge, the overall trends are only weakly dependent on β and γ, at least for 1 ≲ γ ≲ 20.

Figure 11.

Figure 11. Density nrms vs. diameter D. (a) Models with (β, γ)= (2,1), (2,5), (3,10), and (5,20). Results shown were calculated for T4 = 0.94, $\langle h\nu \rangle {\rm_i}=18 \,{\rm eV}$. Solid lines show models with pedge fixed, and Q0 varying from 1048 s−1 to 1054 s−1. Broken lines show models with Q0 fixed, and pedge/k varying from 104 cm−3 K to 1011 cm−3 K. (b) Model grid for β = 3 and γ = 10 together with observed values are shown for various samples of Galactic and extragalactic H ii regions. Cyan open triangles: Kennicutt (1984). Blue diamonds: Churchwell & Goss (1999). Green crosses: Garay & Lizano (1999). Cyan crosses: Kim & Koo (2001). Black open stars: Martín-Hernández et al. (2005). Red solid triangles: radio sample from Hunt & Hirashita (2009). Red open circles: Hubble Space Telescope sample from Hunt & Hirashita (2009).

Standard image High-resolution image

Figure 11(b) shows the model grid for β = 3 and γ = 5 together with observed values of D and nrms from a number of different studies. It appears that observed H ii regions—ranging from H ii regions ionized by one or at most a few O stars (Q0 < 1050 s−1) to "super star clusters" powered by up to 103–105 O stars (Q0 = 1052–1054 s−1)—can be accommodated by the present static equilibrium models with external pressures in the range 104p/k ≲ 1010.3 cm−3 K. Note that for diameters D ≳ 102 pc, the assumption of static equilibrium is unlikely to be justified, because the sound-crossing time (D/2)/15 km s−1 ≳ 3 Myr becomes longer than the lifetimes of high-mass stars.

The fact that some H ii region samples (e.g., Garay et al. 1993; Kim & Koo 2001) seem to obey an nrmsD−1 relationship appears to be an artifact of the sample selection. We see in Figure 11(b) that the overall sample of H ii regions does not have a single nrms versus D relationship. But the observations appear to be generally consistent with the current models of dusty H ii regions.

5.3. Cavities in H ii Regions: N49

Even without dust present, radiation pressure from photoelectric absorption by H and He can alter the density profile in a static H ii region, lowering the central density and enhancing the density near the edge of the ionized region (see Figure 2(a)). As seen in Figure 4(a), for large values of Q0nrms, the surface brightness profile can be noticeably flattened. If dust is assumed to be present, with properties typical of the dust in diffuse clouds, the equilibrium density profile changes dramatically, with a central cavity surrounded by a high-pressure shell of ionized gas pushed out by radiation pressure. In real H ii regions, fast stellar winds will also act to inflate a low-density cavity, or "bubble," near the star; the observed density profile will be the combined result of the stellar wind bubble and the effects of radiation pressure.

The GLIMPSE survey (Churchwell et al. 2009) has discovered and cataloged numerous interstellar "bubbles." An example is N49 (Watson et al. 2008), with a ring of free–free continuum emission at 20 cm, surrounded by a ring of 8 μm PAH emission. An O6.5V star is located near the center of the N49 ring. The image is nearly circularly symmetric, with only a modest asymmetry that could be due to the motion of the star relative to the gas. The 20 cm image has a ring-peak-to-center intensity ratio I(peak)/I(center) ≈ 2.

Is the density profile in N49 consistent with what is expected for radiation pressure acting on dust? From the 2.89 Jy flux from N49 at λ = 20 cm (Helfand et al. 2006) and distance 5.7 ± 0.6 kpc (Churchwell et al. 2006), the stellar source has Q0,49 ≈ (0.78 ± 0.16)/fion. If fion ≈ 0.6, then Q0,49 ≈ (1.3 ± 0.3). The H ii region, with radius (0.018 ± 0.02) deg, has nrms ≈ 197 ± 63 cm−3. Hence, Q0,49nrms ≈ 260 cm−3. If σd,−21 = 1, then τd0 ≈ 1.3. From Figure 6(a), we confirm that fion ≈ 0.6 for τd0 ≈ 1.3.

Figure 5(d) shows that an H ii region with τd0 = 1.3 is expected to have a central minimum in the EM, but with I(peak)/I(center) ≈ 1.3 for β = 3 and γ = 10, whereas the observed I(peak)/I(center) ≈ 2. The central cavity in N49 is therefore significantly larger than would be expected based on radiation pressure alone. While the effects of radiation pressure are not negligible in N49, the observed cavity must be the result of the combined effects of radiation pressure and a dynamically important stellar wind (which is of course not unexpected for an O6.5V star).

5.4. Lyα

The original ionizing photon deposits a radial momentum $h\nu {\rm_i}/c$ at the point where it is absorbed by either a neutral atom or a dust grain. A fraction (1 − fion) of the ionizing photons are absorbed by dust; this energy is re-radiated isotropically, with no additional force exerted on the emitting material. Because the infrared optical depth within the H ii region is small, the infrared emission escapes freely, with no dynamical effect within the H ii region.

A fraction fion of the ionizing energy is absorbed by the gas. Subsequent radiative recombination and radiative cooling converts this energy to photons, but the isotropic emission process itself involves no net momentum transfer to the gas. We have seen above that the H ii can have a center-to-edge dust optical depth τ(R) ≈ 1.6 for τd0 ≳ 5, or Q0,49nrms ≳ 102 cm−3 (cf. Figure 5(c) with β = 3 and γ = 10). This optical depth applies to the hν>13.6 eV ionizing radiation; the center-to-edge optical depth for the hν < 3.4 eV Balmer lines and collisionally excited cooling lines emitted by the ionized gas will be significantly smaller, and much of this radiation will escape dust absorption or scattering within the H ii region. That which is absorbed or scattered will exert a force on the dust at that point only to the extent that the diffuse radiation field is anisotropic. We conclude that momentum deposition from the Balmer lines and collisionally excited cooling lines within the ionized zone will be small compared to the momentum deposited by stellar photons.

Lyα is a special case. At low densities (n ≪ 103 cm−3), ∼70% of Case B recombinations result in emission of a Lyα photon, increasing to >95% for n>105 cm−3 as a result of collisionally induced 2s → 2p transitions (Brown & Mathews 1970). After being emitted isotropically, the photon may scatter many times before either escaping from the H ii or being absorbed by dust. Most of the scatterings take place near the point of emission, while the photon frequency is still close to line center. On average, the net radial momentum transfer per emitted photon will likely be dominated by the last scattering event before the photon escapes from the H ii region, or by the dust absorption event if it does not. At a given point in the nebula, the incident photons involved in these final events will be only moderately anisotropic. Since there is less than one Lyα photon created per Case B recombination, the total radial momentum deposited by these final events will be a small fraction of the radial momentum of the original ionizing photons. Henney & Arthur (1998) estimate that dust limits the Lyα radiation pressure to ∼6% of the gas pressure. We conclude that Lyα has only a minor effect on the density profile within the ionized zone.

5.5. H ii Region Expansion

ii regions arise when massive stars begin to emit ionizing radiation. The development of the H ii region over time depends on the growth of the ionizing output from the central star, and the expansion of the initially high-pressure ionizing gas. Many authors (e.g., Kahn 1954; Spitzer 1978) have discussed the development of an H ii region in gas that is initially neutral and uniform. If the ionizing output from the star turns on suddenly, the ionization front is initially "strong R-type," propagating supersonically without affecting the density of the gas, slowing until it becomes "R-critical," at which point it makes a transition to "D-type," with the ionization front now preceded by a shock wave producing a dense (expanding) shell of neutral gas bounding the ionized region.

While the front is R-type, the gas density and pressure are essentially uniform within the ionized zone. When the front becomes D-type, a rarefaction wave propagates inward from the ionization front, but the gas pressure (if radiation pressure effects are not important) remains relatively uniform within the ionized region, because the motions in the ionized gas are subsonic.

When radiation pressure effects are included, the instantaneous density profile interior to the ionization front is expected to be similar to the profile calculated for the static equilibria studied here. Let Vi be the velocity of the ionization front relative to the star. When the ionization front is weak D-type, the velocity of the ionization front relative to the ionized gas just inside the ionization front is ${\sim} 0.5 V{\rm_i}$ (Spitzer 1978). Given the small dust drift velocities vd,r near the ionization front (i.e., τ(r) → τ(R) in Figure 8), dust is unable to drift outward across the ionization front as long as the ionization front is propagating outward with a speed (relative to the ionized gas) $V{\rm_i}\gtrsim 0.1 \,{\rm km\, s}^{-1}$.

6. SUMMARY

  • 1.  
    Dusty H ii regions in static equilibrium consist of a three-parameter family of similarity solutions, parameterized by the parameters β, γ, and a third parameter, which can be taken to be Q0,49nrms or τd0 (see Equation (16)). The parameter β (Equation (6)) characterizes the relative importance of hν < 13.6 eV photons and γ (Equation (7)) characterizes the dust opacity. A fourth parameter—e.g., the value of nrms or Q0,49—determines the overall size and density of the H ii region.
  • 2.  
    Radiation pressure acting on both gas and dust can strongly affect the structure of H ii regions. For dust characteristic of the diffuse ISM of the Milky Way, static H ii regions with Q0,49nrms ≲ 102 cm−3 will have nearly uniform density; but when Q0,49nrms ≫ 102 cm−3, radiation pressure acts to concentrate the gas in a spherical shell.
  • 3.  
    For given β and γ, the importance of radiation pressure is determined mainly by the parameter τd0 (see Equation (16)). When τd0 ≳ 1, radiation pressure will produce a noticeable central cavity.
  • 4.  
    If the dust-to-gas ratio is similar to the value in the Milky Way, then compression of the ionized gas into a shell limits the characteristic ionization parameter: U1/2 ≲ 0.01, even for Q0nrms ≫ 1 (see Figure 3).
  • 5.  
    For τd0 ≳ 1, compression of the gas and dust into an ionized shell leads to a substantial increase (compared to the estimate by Petrosian et al. 1972) in the fraction fion of hν>13.6 eV photons that actually ionize H, relative to what would have been estimated for a uniform-density H ii region, as shown in Figure 6. Equation (21) allows fion to be estimated for given Q0nrms, β, and γ. Galactic H ii regions appear to have values of fion consistent with the present results for H ii regions with radiation pressure (see Figure 10).
  • 6.  
    Interstellar bubbles surrounding O stars are the result of the combined effects of radiation pressure and stellar winds. For the N49 bubble, as an example, the observed ring-like free–free emission profile is more strongly peaked than would be expected from radiation pressure alone, implying that a fast stellar wind must be present to help create the low-density central cavity.
  • 7.  
    For static H ii regions, dust drift would be important on timescales ≲1 Myr for Q0,49nrms ≳ 103 cm−3. Real H ii regions are not static, and the dust will not drift out of the ionized gas because the ionization front will generally be propagating (relative to the ionized gas just inside the ionization front) faster than the dust drift speed ≲1 km s−1 (see Figure 8).

I am grateful to Bob Benjamin and Leslie Hunt for helpful discussions, to R. H. Lupton for making available the SM graphics package, and to the anonymous referee for suggestions that improved the paper. This research made use of NASA's Astrophysics Data System Service, and was supported in part by NASA through JPL contract 1329088, and in part by NSF grant AST 1008570.

Footnotes

  • For γ>0, u ∝ exp [(β + 1)γ/y] → as y → 0, and the integration must start at some small y>0.

  • Some authors (e.g., Dopita et al. 2006) use the volume-averaged ionization parameter 〈UV. For a uniform-density dustless H ii region, 〈UV = (81/256π)1/32/3B/c)(Q0nrms)1/3 = 2.83 U1/2.

  • This solution is physically relevant if the drift speed began with s ≲ 0.9 and increased with time.

Please wait… references are loading.
10.1088/0004-637X/732/2/100