Articles

EFFECTS OF MAGNETIC FIELD AND FAR-ULTRAVIOLET RADIATION ON THE STRUCTURES OF BRIGHT-RIMMED CLOUDS

, , , and

Published 2013 March 6 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Kazutaka Motoyama et al 2013 ApJ 766 50 DOI 10.1088/0004-637X/766/1/50

0004-637X/766/1/50

ABSTRACT

The bright-rimmed cloud SFO 22 was observed with the 45 m telescope of Nobeyama Radio Observatory in the 12CO (J = 1–0), 13CO (J = 1–0), and C18O (J = 1–0) lines, where well-developed head–tail structure and small line widths were found. Such features were predicted by radiation-driven implosion models, suggesting that SFO 22 may be in a quasi-stationary equilibrium state. We compare the observed properties with those from numerical models of a photoevaporating cloud, which include effects of magnetic pressure and heating due to strong far-ultraviolet (FUV) radiation from an exciting star. The magnetic pressure may play a more important role in the density structures of bright-rimmed clouds than the thermal pressure that is enhanced by the FUV radiation. The FUV radiation can heat the cloud surface to near 30 K; however, its effect is not enough to reproduce the observed density structure of SFO 22. An initial magnetic field of 5 μG in our numerical models produces the best agreement with the observations, and its direction can affect the structures of bright-rimmed clouds.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Bright-rimmed clouds (BRCs) are cometary molecular clouds found at the edge of H ii regions. These clouds have bright rims on the side facing the exciting star and extended tails on the other side. Since their head–tail morphologies suggest that BRCs are interacting with the radiation or stellar wind from an exciting star, BRCs are considered to be potential sites of triggered star formation. The gradients of age spread in young stars along the axes of BRCs indicate that their formation may have been sequentially triggered by shock waves (Sugitani et al. 1995; Getman et al. 2007; Ikeda et al. 2008; Getman et al. 2009).

Radiation-driven implosion models are often considered for the formation and evolution of BRCs and their triggered origins. Strong UV radiation from nearby massive stars can photoionize and photoevaporate the surfaces of surrounding molecular clouds, and whose effects have been studied by various groups. Bertoldi (1989) developed an approximate analytical solution for the evolution of molecular cloud compressed by radiation-driven implosion. Lefloch & Lazareff (1994) investigated a radiation-driven implosion model using hydrodynamic simulations. Recent hydrodynamic simulations of radiation-driven implosion include effects of physics such as self-gravity of the gas (Kessel-Deynet & Burkert 2003; Miao et al. 2006), diffuse radiation field (Haworth & Harries 2012), and turbulence in molecular clouds (Gritschneder et al. 2009). Motoyama et al. (2007) demonstrated that radiation-driven implosion can enhance accretion rates enough to account for the high luminosities of young stellar objects observed in the BRCs (Sugitani et al. 1989). Typical shock speed of a few km s−1 in this study is consistent with that estimated from observations of age gradients of young stars in and around BRCs (Getman et al. 2007, 2009).

Radiation-MHD studies suggest the possibility of altering evolution of photoionized cloud by the presence of a magnetic field. Henney et al. (2009) carried out the first three-dimensional radiation-MHD simulations of photoionization of a magnetized molecular globule under ultraviolet radiation. They reported the photoevaporating globule will evolve into a more flatten shape compared with the non-magnetic case when the cloud initially has a strong magnetic field (that is 100 times the thermal pressure) perpendicular to UV radiation field. Mackey & Lim (2011) also showed that a strong magnetic field has significant influence on the dynamics of the photoionization process. Measuring magnetic field strengths in BRCs observationally has been difficult, and magnetic field effects can only be inferred indirectly by comparing the observed density structures and kinematics with those obtained by theoretical models.

Strong far-ultraviolet (FUV) radiation from an exciting star may also influence the evolution of a photoionized cloud. As shown in many studies of photon dominated regions (e.g., Tielens & Hollenbach 1985; Hollenbach et al. 1991), the FUV radiation from a massive star can heat molecular clouds through photoelectric heating and photodissociation of important coolants such as carbon monoxide. Temperature of the cloud heated by FUV radiation ranges from a few tens up to a few hundred kelvin, strongly dependent on the intensity of the FUV radiation and the density of cloud. High thermal pressure enhanced by FUV radiation may affect the evolution of BRCs. However, there have been limited theoretical works that include heating due to FUV radiation (Miao et al. 2006; Henney et al. 2009). These effects should be included in theoretical models for reliable comparisons with observations.

The purpose of this study is to investigate how magnetic field and FUV radiation affect the evolution of BRCs through the comparisons between observations and numerical models. An evolutionary scenario of BRCs due to radiation-driven implosion has an initial implosion phase followed by a quasi-stationary equilibrium phase. In this paper, we focus on the quasi-stationary equilibrium phase, and implosion phase will be investigated in a subsequent paper. The BRC SFO 22 was observed with the 45 m telescope of Nobeyama Radio Observatory, and the results were compared with numerical models of photoionized clouds with the effects of magnetic field and FUV radiation. The layout of this paper is as follows. In Section 2, the details of observations are described. In Section 3, a description of our numerical models is given. Sections 4 and 5 give the results and discussions. In Section 6, we summarize our main conclusions.

2. OBSERVATIONS AND ANALYSIS

2.1. Observed Bright-rimmed Cloud

BRC SFO 22 is located at the eastern edge of H ii region s281, and was selected from BRC catalog of Sugitani et al. (1991). Figure 1(a) shows the entire image of the region s281, and Figure 1(b) gives the close-up view of SFO 22 enlarged from Figure 1(a). This H ii region is ionized by θ1Ori, marked by a cross in Figure 1(a). The spectral type of the primary exciting star in θ1Ori is O7V, and has a projected distance of 6.5 pc to SFO 22 (Morgan et al. 2004). S281 (Blaauw 1964) is 460 pc away from us.

Figure 1.

Figure 1. (a) DSS red image of the H ii region s281. The cross labels the position of the exciting star. (b) DSS red image of bright-rimmed cloud SFO 22. The arrow indicates the direction to the exciting star.

Standard image High-resolution image

Sugitani et al. (1991) classified the BRCs into three types in order of increasing degree of rim curvature: types A, B, and C. Based on the radiation-driven implosion model, the shapes of cloud rims reflect the evolutionary stages. The type A BRCs are still undergoing compression by the shock waves generated by ionization, and the type B and type C BRCs are approaching or have already reached the phase of quasi-stationary equilibria. SFO 22 has a well-developed head–tail structure along the line to the exciting star and it is classified as type B. Although the head part of the cloud contains the IRAS point source 05359-0515, ammonia rotational inversion lines were not detected toward this IRAS point source (Morgan et al. 2010). Since ammonia lines trace dense gas associated with protostellar cores, the non-detection of ammonia lines has been interpreted as no star forming activities in SFO 22.

2.2. Observations

Observations were carried out using the 45 m telescope of Nobeyama Radio Observatory in 2005 January and April. We observed 12CO (J = 1–0) at 115.271203 GHz, 13CO (J = 1–0) at 110.201370 GHz, and C18O (J = 1–0) at 109.782182 GHz. The half-power beam width of the telescope and main-beam efficiency at 115 GHz were 15'' and ηMB = 0.4, respectively. We used the 5 × 5 beam focal plane array receiver "BEARS" whose beam separation is 41farcs1. As receiver backends we used a 1024 channel digital autocorrelator with a 31.25 kHz frequency resolution. The corresponding velocity resolution is 81.5 m s−1 at 115 GHz. The typical system noise temperature was 300–450 K depending on the atmospheric conditions. The intensity scale of the spectra was calibrated by the chopper wheel method. The corrected antenna temperature $T^*_A$ is converted into main beam brightness temperature using the relation of $T_B = T^*_A /\eta _{{\rm mb}}$.

We observed SFO 22 with a grid spacing of 20farcs55, which is half of the beam separation of BEARS, in the line of 12CO (J = 1–0). The mapped area was 390'' × 390''. Dense regions of clouds were observed with finer grid spacing of 10farcs3 in the lines of 13CO (J = 1–0) and C18O (J = 1–0). The mapped area for these lines were 195'' × 195'' and 154'' × 154'', respectively. The pointing accuracy of the antenna was checked and corrected every 1.5–2 hr using SiO maser emission from Ori KL, and its typical error was less than 5''. All observations were carried out by position switching mode. The data were reduced by using the software package NewStar provided by Nobeyama Radio Observatory.

2.3. Observational Results and Analysis

Figure 2 shows the velocity-integrated intensity maps of SFO 22. The reference center of the map is the peak position of C18O (J = 1–0) emission at R.A. (2000) = $5\mathrm{^h} 38\mathrm{^m} 21\buildrel{\mathrm{m}}\over{.}6$, decl. (2000) = −5°13'37farcs8. Emission from 12CO (J = 1–0) coincides with the optical images. The cometary morphology is clearly shown. On the contrary, the C18O (J = 1–0) emission is very weak and detected only at a few points.

Figure 2.

Figure 2. Integrated intensity maps of SFO 22. The 12CO (J = 1–0) map (top) has a lowest contour of 1.39 K km s−1 (3σ) and contour intervals of 1.39 K km s−1 (3σ). The 13CO (J = 1–0) map (middle) has a lowest contour of 0.315 K km s−1 (3σ) and contour intervals of 0.525 K km s−1 (5σ). The C18O (J = 1–0) map (bottom) has a lowest contour of 0.131 K km s−1 (3σ) and contour intervals of 0.0435 K km s−1 (1σ).

Standard image High-resolution image

Figure 3 shows the observed line spectra toward the peak position of C18O (J = 1–0) and offset positions along declination. Assuming that 12CO (J = 1–0) emission is optically thick at the peak of C18O (J = 1–0) emission, we can calculate the excitation temperature as

Equation (1)

where TB(12CO) is the brightness temperature of 12CO (J = 1–0) at the peak of C18O (J = 1–0). The excitation temperature of SFO 22 is found to be 27.1 ± 1.8 K. Figure 4 shows the position–velocity diagrams of 12CO (J = 1–0), 13CO (J = 1–0), and C18O (J = 1–0) along declination through the peak position of C18O (J = 1–0). The line widths of 12CO (J = 1–0), 13CO (J = 1–0), and C18O (J = 1–0) are roughly ∼2.0 km s−1, ∼1.5 km s−1, and ∼1.0 km s−1, respectively. Line widths of SFO 22 are relatively narrow compared to other BRCs. In the millimeter and submillimeter molecular line survey of BRCs by De Vries et al. (2002), many BRCs have CO line widths of ≳5 km s−1. These large line widths of BRCs may be attributed to large velocity dispersion due to turbulence and outflow activities. The observed narrow line widths of SFO 22 could suggest that the influence from the photoionization-induced shocks may have already disappeared, and SFO 22 has now reached the phase of quasi-stationary equilibrium predicted in the radiation implosion model.

Figure 3.

Figure 3. Some selected 12CO (J = 1–0), 13CO (J = 1–0), and C18O (J = 1–0) spectra observed toward the SFO 22 along declination through the peak of C18O (J = 1–0).

Standard image High-resolution image
Figure 4.

Figure 4. Position–velocity diagrams of 12CO (J = 1–0) (left), 13CO (J = 1–0) (center), and C18O (J = 1–0) (right) along declination through the peak positions of C18O (J = 1–0) emission. The diagram for 12CO (J = 1–0) has lowest contour of 1.53 K (3σ) and contour interval of 1.53 K (3σ). The diagram for 13CO (J = 1–0) has lowest contour of 0.276 K (3σ) and contour interval of 0.92 K (10σ). The diagram for C18O (J = 1–0) has lowest contour of 0.125 K (3σ) and contour interval of 0.125 K km (3σ). The vertical dotted lines indicate the systemic velocity of 10.4 km s−1.

Standard image High-resolution image

Figure 5 shows column density profiles of SFO 22 along the declination, which is almost parallel to the direction to the exciting star. We derive the column density distribution of the cloud using data from this observation. Assuming that the excitation temperatures of the 13CO and C18O lines are the same as the 12CO line, we can calculate the optical depth of 13CO and C18O using

Equation (2)

and

Equation (3)

respectively. The column densities of 13CO and C18O molecules can be derived as

Equation (4)

and

Equation (5)

where Δv13 and Δv18 are the line widths of the 13CO (J = 1–0) and C18O (J = 1–0) emission, respectively. The column density of 13CO is converted to the column density of H2 by assuming an abundance ratio of N(H2)/N(13CO) = 5.0 × 105 (Dickman 1978). Under the environment where molecular clouds are illuminated by strong UV radiation, abundance ratio of N(H2)/N(C18O) are thought to be affected by selective destruction by UV radiation (Glassgold et al. 1985). We follow Niwa et al. (2009) to obtain the column density of H2 from the column density of C18O. The column densities of 13CO and C18O can be formulated by least-squares fitting as

Equation (6)

An abundance ratio of N(H2)/N(C18O) = 1.28 × 107 can be derived. This value is slightly larger than the standard value of 6.0 × 106 for molecular clouds not associated with H ii regions (Frerking et al. 1982). Figure 5 shows that the column density profiles obtained here are nearly flat with the column density of ∼1022 cm−2. Adopting a distance of 460 pc to SFO 22, the total masses traced by 13CO emission and C18O emission are M13 = 12.0 M and M18 = 5.3 M, respectively.

Figure 5.

Figure 5. Column density profiles of SFO 22 along the declination. The open circles (blue in the online journal) and filled circles (colored red in the online version) label the column densities derived from 13CO (J = 1–0) emission and C18O (J = 1–0) emission, respectively.

Standard image High-resolution image

3. MODEL DESCRIPTION

Our numerical model is based on the analytic model of Bertoldi & McKee (1990) for a photoevaporating cloud in the quasi-stationary equilibrium state with a polytropic equation of state. The polytropic gas is a good approximation for the cases where either gas pressure or magnetic pressure is dominant. In an actual bright-rimed cloud, these two quantities may have been comparable before the compression by radiation-driven implosion alter the state, and FUV radiation will affect the thermal structure through photoelectric heating and photodissociation of molecular coolants. Here we adopt a more realistic model with the inclusion of thermal pressure and magnetic pressure explicitly and the heating due to FUV radiation.

3.1. Density Structure

Total pressure of the gas is expressed as

Equation (7)

where, Pth, Pmag, cs, ρ, γ, and B are the thermal pressure, the magnetic pressure, the sound speed, the density of the cloud, the ratio of specific heats, and the magnetic field strength, respectively. A γ = 5/3 is adopted, which is appropriate for the molecular clouds because the molecular hydrogen behaves like a monoatomic gas at temperature ≲100 K. We assume an ideal gas, so that the sound speed cs is related to the gas temperature T as

Equation (8)

where kB and μ are the Boltzmann constant and mean mass per nucleus in units of the hydrogen mass mH = 1.67 × 10−24 g and μ = 1.15, respectively. For simplicity, the effects of the magnetic field are approximated through

Equation (9)

where B0 and ρ0 are magnetic field strength and density of the cloud before undergoing compression by radiation-driven implosion. The exponent α, which ranges from 0 to 1, is a parameter representing how much the magnetic field is trapped in the gas during the compression. Figure 6 shows schematic drawings of two extreme cases by assuming that compression of a cloud perpendicular to the direction of radiation is small. If the cloud is compressed along the magnetic field, the magnetic field strength will hardly change during compression, i.e., α ≃ 0. On the other hand, if the cloud is compressed perpendicular to the magnetic field, the strength of the field increases as B∝ρ during compression, i.e., α ≃ 1. The value of α depends on the initial configuration of the magnetic field and the shape of cloud. We leave α as an open parameter as it is hard to determine an accurate value without launching MHD simulations. We also neglect the diffusion of the magnetic field due to the longer timescale of ambipolar diffusion compared to the dynamical timescale of the radiation-driven implosion.

Figure 6.

Figure 6. Schematic figures of compression of magnetized cloud due to radiation-driven implosion for two extreme cases. The UV radiation propagates downward, and the cloud is compressed only along this direction in both cases. (Left) The magnetic field is parallel to the direction of the UV radiation. (Right) The magnetic field is perpendicular to the direction of the UV radiation.

Standard image High-resolution image

We calculate the structure of the cloud in axisymmetry along the line to the exciting star. We also assumed that the distance from the cloud to the exciting star is larger than size of the cloud, so that the UV radiation field can be treated as planar. Figure 7 illustrates the coordinates system we use in this paper. UV radiation propagates downward parallel to the z-axis, and the ionized gas evaporates off the cloud surface with angle θ to z-axis. Bertoldi & McKee (1990) showed that the position of the cloud surface can be approximated as

Equation (10)

where Rc and a are the curvature radius at z = 0 and the cloud width parameter defined in Bertoldi & McKee (1990), respectively. The method used to determine a and Rc is described in Section 3.3.

Figure 7.

Figure 7. Coordinate system used in our numerical models. The exciting star is assumed far above. The ionizing UV radiation is parallel to the symmetric axis. The flow of evaporation streams along the surface normal. The angle between the surface normal and UV radiation is denoted by θ.

Standard image High-resolution image

In this study, we assume that all parts of the cloud are subject to the same acceleration through the rocket effect of the photoevaporation flow, as the cloud is in a quasi-stationary equilibrium state. The equation of hydrostatic equilibrium for the cloud is

Equation (11)

where g is the acceleration of the cloud. With Equations (7) and (9), this equation can be rewritten as

Equation (12)

The density distribution inside the cloud is determined by solving this equation with the appropriate boundary conditions.

The boundary conditions at cloud surface are determined using jump conditions for D-critical ionization front. Therefore, the pressure at the cloud surface is given by

Equation (13)

where ci = 10 km s−1 and FUV(r) are the sound speed of the ionized gas and the ionization photon flux reaching the ionization front, respectively, and the width of the ionization front is negligible. Since some part of the incident ionizing UV photons is consumed by recombined hydrogen in the photoevaporation flow, the ionizing UV photon flux arriving at the cloud surface is written as

Equation (14)

where Fi, αB = 2.7 × 10−13 cm3 s−1, ne, and np are the incident ionizing photon flux, the hydrogen electronic recombination coefficient into the excited state, the electron number density, and the proton number density, respectively. Following Bertoldi & McKee (1990), we introduce a dimensionless parameter ω that represents the effective fractional thickness of the recombination layer. Equation (14) can be rewritten as

Equation (15)

where nII is the hydrogen number density just behind the ionization front. If we assume that the ionized gas streams away from the ionization front with the velocity of the speed of sound, and the stream line is normal to the cloud surface, we can derive the form

Equation (16)

where R(r') is the curvature radius of the cloud surface at r = r'. We adopt ω for a spherical cloud with radius Rc at the cloud tip:

Equation (17)

where q and ψ are the ratio of incident ionizing photon flux Fi to ionizing photon flux reaching the ionization front FUV and the photoevaporation parameter, respectively. The photoevaporation parameter is defined as

Equation (18)

The value of ψ in our numerical models ranges from 74 to 104 depending on the model parameters. Spitzer (1978) derived an analytic estimation of q as

Equation (19)

The density and pressure distributions of the cloud are characterized by scale height defined as hc = Ptot(z = 0)/gρ(z = 0). If ψ ≫ 1, which is the case for SFO 22, hc is related to Rc as Rc = 0.5hc.

3.2. Thermal and Chemical Model

We solve the reaction networks for the species of H2, CO, C+, O, and the electron e. We adopt the simplified reaction model as described in Nelson & Langer (1997) to determine the abundance of CO molecules. In this reaction model, C+ is directly converted to CO without accounting explicitly for the intermediate reactions. The formation rate of CO molecules is expressed as

Equation (20)

where n(C+) and n(H2) are the number densities of C+ and H2, respectively. The coefficient β is defined as

Equation (21)

where X(O) and $D_{\mathrm{CH_x}}$ are the fractional abundance of oxygen and the total photodissociation rate of both CH and CH2. This total photodissociation rate is written as

Equation (22)

where G0 and τUV are the intensity of the incident FUV radiation in terms of the Habing interstellar radiation field (Habing 1968) and the optical depth in UV range, respectively. The optical depth τUV is related to the visual extinction Av by τUV = 2.5Av. We adopt the conversion factor between the visual extinction and the total hydrogen column density as $X_{A_V} = A_V / N_{\rm H} = 6.3 \times 10^{-22}$ mag cm2 in this paper. The photodissociation rate of CO by the FUV radiation is written as

Equation (23)

where n(CO) is the number density of CO. The abundance of CO is calculated by solving the equation of formation and dissociation balance

Equation (24)

Abundances of other species are calculated as follows. The cloud is assumed to be composed of molecular hydrogen, so that the number density of H2 is written as

Equation (25)

where n is the number density of hydrogen nuclei and related to the mass density as

Equation (26)

We assume that carbon exists in ionized form in the cloud as carbon is easily photoionized by FUV radiation owing to its lower ionization energy (11.2 eV) compared to that of hydrogen. Although this assumption may cause an overestimation of the cooling rate through C+ in the deep inner region of the cloud where FUV is strongly attenuated and rotational line emission from CO is a more important cooling process. Our assumption does not affect the numerical results, however. The number density of C+ is calculated by

Equation (27)

where the elemental abundance of carbon is taken to be $X(\mathrm{C_{{\rm tot}}}) = 10^{-4}$. We assume that oxygen exists in atomic form in the cloud, because its ionization energy (13.6 eV) is similar to that of hydrogen. The number density of oxygen is calculated by

Equation (28)

where the elemental abundance of oxygen is taken to be $X(\mathrm{O_{{\rm tot}}}) = 2.0 \times 10^{-4}$. Constant electron fraction ne/n = 10−7 is also assumed to calculate the number density of the electron.

For the heating processes in the cloud, we consider photoelectric heating and cosmic ray heating. The photoelectric heating rate is (Bakes & Tielens 1994)

Equation (29)

where epsilon is the photoelectric heating efficiency

Equation (30)

Cosmic ray heating becomes the dominant heating process in the inner region where FUV radiation does not penetrate. Cosmic ray heating rate is given by

Equation (31)

where ζp(H2) is the primary cosmic ray ionization rate of H2 and ΔQcr is the energy deposited as heat as a result of this ionization. We adopt values of ζp(H2) = 7.0 × 10−17 s−1 (van Dishoeck & Black 1986) and ΔQcr = 20 eV (Goldsmith & Langer 1978).

For the cooling processes in the cloud, we consider radiation from the CO rotational transitions, collisionally excited line emission from C+ and O, and collisional heat transfer between gas and dust. The cooling rate due to CO, ΛCO, is taken from the tabulated cooling function computed by Neufeld et al. (1995) for T ⩽ 100 K and Neufeld & Kaufman (1993) for T > 100 K. The cooling rates due to the collisional excitation of C+ and O are taken from Nelson & Langer (1997). The cooling rate due to C+ is

Equation (32)

where the critical number density is taken to be ncrit = 3 × 103 cm−3. The cooling rate due to O is

Equation (33)

The cooling rate due to the dust grain is (Hollenbach & McKee 1989)

Equation (34)

where amin and Tdust are the minimum radius of grains and the dust temperature, respectively. We used amin = 100 Å to calculate the cooling rate by dust grains. The dust temperature is calculated following the method by Hollenbach et al. (1991).

3.3. An Iterative Procedure for Obtaining Numerical Solution

An iterative procedure is used to achieve the density distribution of the cloud. We use a uniform grid of 200 (radial) × 1000 (axial) cells. The grid spacing is 6.5 × 10−4 pc. For presentation, we symmetrize the results obtained on a grid of 400 × 1000 cells. In order to fit the width of the cloud with that of SFO 22, the cloud width parameter a is calculated by substituting r = 0.13 pc and z = −0.6 pc into Equation (10). The procedure is as follows. (1) Equation (12) is numerically solved by the Runge–Kutta method using current temperature and sound speed. (2) The optical depth to the cloud surface is calculated. (3) The chemical reaction network is solved to determine the abundances of included species. (4) The temperature and sound speed are updated using new chemical abundances. The thermal equilibrium is assumed to determine the temperature of the cloud:

Equation (35)

These processes are repeated until the converged solution is achieved. We have confirmed that relative errors between the two consecutive steps are smaller than 10−2 for all models in this paper.

Once the curvature radius of the cloud Rc is given, we can determine the position of the cloud surface from Equation (10) and calculate the structure of the cloud using the iterative procedure described above. Since we assume that the cloud is in an equilibrium state, we adopt Rc, which minimizes the pressure gradient along the r-direction. We change Rc from 0.02 pc to 0.11 pc in increments of 0.001 pc, and calculate the structure of the cloud for each Rc. Then, we integrate the deviation of the total pressure Ptot(r, z) from the average total pressure over all cells at the same z coordinates Pave(z),

Equation (36)

We adopt the structure of the cloud for which this value is minimized.

4. RESULTS

We present the results of our numerical models in this section. Model parameters of the run are summarized in Table 1. To compare with our observational results, intensities of the incident ionizing UV and FUV radiation were determined to correspond to those expected in the region where SFO 22 is located. Since the spectral type of the primary exciting star of SFO 22 is O7V, we adopted the values of log SUV = 48.76 s−1 and log SFUV = 48.76 s−1 as the UV and FUV photon luminosities (Vacca et al. 1996; Diaz-Miller et al. 1998). These luminosities and the distance from the exciting star to SFO 22 of 6.5 pc give the incident ionizing UV flux of Fi = 1.192 × 109 cm−2 s−1 and incident FUV flux of G0 = 94 in terms of the Habing field FH = 1.21 × 107 cm−2 s−1 (Bertoldi & Draine 1996). We used these values in all models except model A. The number density of the cloud before undergoing compression by radiation-driven implosion was assumed to be n0 = ρ0mH = 103 cm−3 in all models. Some physical values obtained from our results are summarized in Table 2. In Figure 8, column density profiles along the z-axis for all models are plotted with the column densities of SFO 22 derived from our observations.

Figure 8.

Figure 8. Comparisons of the column density profiles along the z-axis with the column densities derived from observations. The top, middle, and bottom panels show the results with α = 0.25, 0.50, and 0.75, respectively. The dashed line (green in the online journal) represents the column density profile of model A. Column densities of SFO 22 are plotted with the same symbols as in Figure 5.

Standard image High-resolution image

Table 1. Model Parameters

Model α B0 G0 Fi
(μG) (cm−2 s−1)
A  ⋅⋅⋅ 0 1.0 1.192 × 109
B  ⋅⋅⋅ 0 94.0 1.192 × 109
C1 0.25 25 94.0 1.192 × 109
C2 0.25 45 94.0 1.192 × 109
C3 0.25 60 94.0 1.192 × 109
C4 0.25 80 94.0 1.192 × 109
D1 0.50 7 94.0 1.192 × 109
D2 0.50 15 94.0 1.192 × 109
D3 0.50 25 94.0 1.192 × 109
D4 0.50 45 94.0 1.192 × 109
E1 0.75 2 94.0 1.192 × 109
E2 0.75 5 94.0 1.192 × 109
E3 0.75 10 94.0 1.192 × 109
E4 0.75 25 94.0 1.192 × 109

Download table as:  ASCIITypeset image

Table 2. Summary of Numerical Results

Model Rc Ttipa $n_{\mathrm{H_2},\max}$b $N_{\mathrm{H_2},\max }$c Bmaxd Mcle
(pc) (K) (cm−3) (cm−2) (μG) (M)
A 7.6 × 10−2 19.5 4.48 × 105 1.26 × 1023  ⋅⋅⋅ 61.3
B 9.4 × 10−2 28.9 4.19 × 105 1.27 × 1023  ⋅⋅⋅ 54.8
C1 7.0 × 10−2 30.3 2.61 × 105 5.65 × 1022 119 24.1
C2 4.1 × 10−2 29.5 1.07 × 105 1.61 × 1022 172 5.59
C3 3.2 × 10−2 27.2 5.35 × 104 6.36 × 1021 192 2.03
C4 2.7 × 10−2 24.1 2.27 × 104 2.15 × 1021 208 0.653
D1 8.2 × 10−2 30.3 1.81 × 105 5.33 × 1022 133 27.7
D2 7.1 × 10−2 27.5 5.75 × 104 1.81 × 1022 161 10.4
D3 6.7 × 10−2 24.2 2.39 × 104 7.17 × 1021 173 4.20
D4 6.5 × 10−2 21.2 7.83 × 103 2.27 × 1021 178 1.35
E1 9.0 × 10−2 30.1 1.42 × 105 5.17 × 1022 139 28.4
E2 8.9 × 10−2 26.9 4.87 × 104 2.05 × 1022 155 13.0
E3 8.8 × 10−2 23.8 2.08 × 104 8.77 × 1021 164 5.81
E4 8.7 × 10−2 20.9 6.35 × 103 2.65 × 1021 168 1.79

Notes. aTemperature at cloud tip. bMaximum value of number density of H2. cMaximum value of column density of H2. dMaximum value of magnetic field strength. eTotal cloud mass.

Download table as:  ASCIITypeset image

4.1. Effects of Strong FUV Radiation and Magnetic Field

In this subsection, we describe the results of the following three typical models A, B, and E2 to show how strong FUV radiation and magnetic field affect structures of clouds. Figure 9 shows the densities, the column densities, and the temperatures of these models. In model A, as a reference, we calculated the structure of the cloud without a magnetic field, assuming the strength of an average interstellar FUV radiation field of G0 = 1. In model B, to see the effects of strong FUV radiation, we calculated the structure of the cloud assuming the strength of the FUV radiation expected in the region where SFO 22 is located, but the magnetic field effects were not included. In model E2, we included not only the effects of strong FUV radiation but also the magnetic field effects. The initial magnetic field strength B0 and α were set to be 5 μG and 0.75, respectively.

Figure 9.

Figure 9. Density, column density, and temperature distributions in models A (left), B (center), and E2 (right).

Standard image High-resolution image

Figure 8 shows that the column density distribution of model A is much different from that of observed cloud. The peak column density of 1.26 × 1023 cm−2 is one order of magnitude higher than that of the observed cloud (∼2 × 1022 cm−2). In addition, the slope of the column density profile is steeper than the observations. Figure 9 shows that the cloud has a nearly constant temperature of 20 K in model A. Since FUV radiation is not significant in this model, effective cooling keeps the dense region relatively cold.

Comparisons of model B to model A show that FUV radiation has little influence on the density and the thermal structures of the cloud. Although the cloud surface is heated near 30 K in model B, this warm surface layer is very thin. High density at the head region prevents CO molecules, which are main coolant of molecular gas, from photodissociation except for a thin surface layer. Isothermal gas is good approximation at inner region of the cloud. Figure 8 shows that the column density of model B is much higher than observations as well as model A. Although FUV radiation reduces the density than model A by factor of a few, it is not enough to reproduce observed low column densities. The shape of the cloud slightly differs from model A. Strong FUV radiation enhances thermal pressure at the head region and makes the curvature radius at the cloud tip larger than that of model A. Moving to the tail side, the differences of the density and the column density from those of model A become larger. As a result, the slope of the column density profile of model B is steeper than that of model A.

Comparisons of model E2 to model A and model B show that the magnetic field reduces the density of the cloud. Figure 9 shows that the density and column density of model E2 at the head region are one order of magnitude lower than those of models A and B. As shown in 4.2, magnetic pressure is dominant at z > −0.4 pc in model E2. Additional support due to magnetic pressure makes the density of the cloud lower than models without a magnetic field. The warm surface layer of model E2 is thicker than that of model B, because lower density at the head region allows FUV radiation to penetrate deeper inside the cloud. Figure 8 shows that the slope of the column density profile of model E2 is flatter than those of models A and B. The column density profile of model E2 shows better agreement with that of the observed cloud than the other two models.

4.2. Dependence on the Density Dependence of Magnetic Field Strength

In this subsection, we present a comparison of three models, C2, D2, and E2, to show how the value of exponent α in Equation (9) affects the results. The values of α and B0 in models C2, D2, and E2 were set to be 0.25 and 45 μG, 0.50, and 15 μG, and 0.75 and 5 μG, respectively. The comparison of these models reveals that the value of α influences the structure of cloud.

As can be seen from Figure 8, although these three models have similar maximum values of column density of ∼2 × 1022 cm−2, the column density profiles of these models are qualitatively different. The model with a smaller value of α has a steeper column density profile at the head region (z > −0.2 pc) and has a flatter column density profile at the tail region (z < −0.3 pc). In model C2, the slope of the column density profile becomes flatter moving to the tail side. The opposite trend is observed in model E2. The slope of the column density profile becomes steeper moving to the tail side. Contrary to models C2 and E2, the slope of the column density profile is nearly constant through the entire region in model D2.

Figure 10 shows the magnetic field strength and plasma beta, which is defined as the ratio of thermal pressure to magnetic pressure, in these models. Although these three models have similar magnetic field strengths of ∼150 μG at the cloud tip, the magnetic field strength decreases more quickly moving to the tail side in the model with the larger value of α. The distribution of the plasma beta in the cloud qualitatively changes whether or not α exceeds 0.5. As can be seen from Equations (7) and (9), the thermal pressure and the magnetic pressure are proportional to ρ and ρ, respectively. When α is larger than 0.5, magnetic pressure increases faster than thermal pressure with increasing density, and magnetic pressure become dominant in the dense region. When α is smaller than 0.5, by contrast, thermal pressure becomes dominant in the dense region. In model C2 in which α = 0.25, the plasma beta is larger than unity in the dense head region and decreases the moving to the tail side. Thus, the thermal pressure is dominant at the head region; the magnetic pressure is dominant at the tail side. In model E2 in which α = 0.75, the opposite trend is observed. The magnetic pressure is dominant at the head region; the thermal pressure is dominant at the tail side. In model D2 in which α = 0.5, the plasma beta is nearly constant in the entire cloud because the magnetic pressure and the thermal pressure increase at the same rate with increasing density.

Figure 10.

Figure 10. Magnetic field strength and ratio of thermal pressure to magnetic pressure in models C2 (left), D2 (center), and E2 (right).

Standard image High-resolution image

4.3. Dependence on Initial Magnetic Field Strength

In this subsection, we describe general trends which arise when the initial magnetic field strength B0 is changed. In Figures 11 and 12, maximum values of the number density and the magnetic field strength are plotted as a function of initial magnetic field strength B0, respectively. As shown in Figures 9 and 10, the number density and the magnetic field strength reach their maximum values near the cloud tip, when the effects of FUV radiation are not significant. We focus on the low plasma regime and consider these maximum values as values at the cloud tips.

Figure 11.

Figure 11. Maximum number densities as function of initial magnetic field strength. The circles (blue in the online journal), the diamonds (red in the online journal), the pluses (green in the online journal), and the asterisk (light blue in the online journal) represent models C1–C4, D1–D4, E1–E4, and B, respectively. The solid lines represent analytic estimates given by Equation (38).

Standard image High-resolution image
Figure 12.

Figure 12. Maximum magnetic field strengths as a function of initial magnetic field strength. The circles (blue in the online journal), the diamonds (red in the online journal), and the pluses (green in the online journal) represent models C1–C4, D1–D4, and E1–E4, respectively. The solid lines represent analytic asymptotic values given by Equation (39).

Standard image High-resolution image

We can obtain the asymptotic behavior of the number density in the low plasma beta regime by neglecting thermal pressure. From Equations (7) and (9), the number density is expressed as

Equation (37)

where we neglect the first term in the right-hand side of Equation (7) because we assume that the magnetic pressure is dominant. The total pressure Ptot at the cloud tip is given by substituting θ = 0 into Equation (13). Therefore, using the relation of FUV = Fi/q, we obtain the number density at the cloud tip as

Equation (38)

This relation shows that the density at the cloud tip is proportional to $B_0^{-1 / \alpha }$ in the low plasma beta regime. The maximum number densities of our numerical models approach lines representing this analytic asymptotic value as B0 increases (see Figure 11). We adopted the minimum value of Rc for models with each α to calculate q in Equation (38), because the curvature radius at the cloud tip Rc decreases with increasing of B0. We used the value of Rc = 0.027 pc for models with α = 0.25, Rc = 0.065 pc for models with α = 0.50, and Rc = 0.087 pc for models with α = 0.75. The estimate value of the maximum number density of SFO 22 is also plotted in this figure. Dividing the maximum value of the observed column density 1.89 × 1022 cm−2 by the cloud width of 0.26 pc gives a rough estimate value of nobs = 2.36 × 104 cm−3. From this figure, it is inferred that SFO 22 had an initial magnetic field strength of several to ∼90 μG, if the value of α is from 0.25 to 0.75.

We can also obtain the asymptotic behavior of the magnetic field strength in the low plasma beta regime. Substituting Equations (26) and (38) into Equation (9) gives the magnetic field strength at the cloud tip as

Equation (39)

This relation shows that the magnetic field strength at the cloud tip is a function of the curvature radius of the cloud at the tip Rc and the incident ionizing photon flux Fi, but is independent of the initial magnetic field strength B0 and the initial number density n0. As the initial magnetic field strength B0 increases, the magnetic field strength at the cloud tip approaches this asymptotic value. Equation (39) gives asymptotic values of 212 μG for models with α = 0.25, 180 μG for models with α = 0.50, and 170 μG for models with α = 0.75. The maximum magnetic field strengths of our numerical models asymptotically approach this value at large B0 (see Figure 12).

5. DISCUSSION

5.1. Comparison between Observations and Numerical Models

We developed a numerical model for a photoevaporating cloud which is in a quasi-stationary equilibrium state, assuming that the pressure gradient of the cloud balances the inertia force caused by acceleration due to the back reaction of the photoevaporation flow. As shown in Section 4.1, the observed density structure of BRC SFO 22 cannot be explained by the reference model (model A) in which either the effects of strong FUV radiation or the effects of magnetic field were not included. The column density of the reference model is one order of magnitude higher than that of the observed cloud. The observed lower column density implies that the pressure of the observed cloud is higher than that of the reference model. There are two possible explanations for this large difference of density structures between the observed cloud and the reference model. One explanation is that heating due to strong FUV radiation from the exciting star warms the cloud, and hence enhanced thermal pressure makes the cloud reach an equilibrium state with lower density than the reference model. Another explanation is that additional pressure due to the magnetic field makes the cloud reach an equilibrium state with lower density than the reference model.

Comparisons between the reference model and numerical model including the effects of strong FUV radiation (model B) show that strong FUV radiation has little influence on the structure of the cloud. Although strong FUV radiation slightly reduces the density and the column density of the cloud, the column density is much higher than that of the observed cloud. The small difference from the reference model suggests that the heating due to the FUV does not affect the structure of the cloud at least in SFO 22. To affect the structure of the cloud, the photoelectric heating needs to be the dominant heating process, i.e., Γpecr > 1. For a rough estimation, let us approximate the efficiency of photoelectric heating epsilon expressed by Equation (30) by a constant value of 4.87 × 10−2. From Equations (29)–(31), and the relation of τUV = 2.5AV, the ratio of the photoelectric heating rate to the cosmic ray heating rate is written as

Equation (40)

Therefore, visual extinction of the cloud along the direction to the exciting star needs to be smaller than the critical value

Equation (41)

Applying this equation to SFO 22 gives the critical value AV, cri as 3.3. The total mass traced by the 13CO emission and width of SFO 22 give a rough estimate of the visual extinction of SFO 22 as

Equation (42)

This gives the value of 15.4. This larger value of AV, SFO22 compared to AV, cri indicates that the heating due to FUV radiation affects only near the cloud surface. This discussion based on a rough estimation is consistent with our numerical results.

Comparisons between the reference model and numerical model including magnetic field effects show that the magnetic field strongly affects the structure of the cloud. From results of models C2, D2, and E2, an initial magnetic field strength of 5–45 μG is required to reproduce the observed column density of ∼2.0 × 1022 cm−2 depending on the value of α. On the other hand, magnetic field strengths in molecular clouds are measured by Zeeman effects. Crutcher et al. (2010) statistically analyzed samples of clouds with Zeeman observation in order to infer the distribution of the total magnetic field strength in the samples. According to their analysis, molecular clouds with the density of ∼103 cm−3 have magnetic field strength of a few to several tens of μG. This coincidence of magnetic field strengths between numerical models and observations shows that the observed column density of SFO 22 is naturally explained by magnetic field effects.

5.2. Magnetic Field Configurations in BRCs

As shown in 4.2, the structure of BRC strongly depends on the value of α in Equation (9). The slope of the column density profile at the head region becomes steeper with decreasing α. The slope of the column density profile at the tail region shows an opposite dependence on α. Since this parameter represents how much the magnetic field is trapped in the gas during the compression, our numerical results imply that the direction of the magnetic field affects the evolution of BRCs. When the magnetic field is parallel to the direction of UV radiation, the value of α is close to 0. On the other hand, when the magnetic field is perpendicular to the direction of UV radiation, the value of α is close to 1. Since the results of model E2, in which α was set to be 0.75, show the best agreement with observations, the magnetic field in SFO 22 is thought to be nearly perpendicular to the UV radiation. However, deviation from observations can be seen at the tail region. The slope of the column density of model E2 is steeper than that of the observed cloud at the tail region (see Figure 8). This difference of column density profiles implies the possibility that the value of α at the tail region is smaller than the head region. Mackey & Lim (2011) performed three-dimensional MHD simulations and found that for weak and medium magnetic field strengths an initially perpendicular field is swept into alignment with the tail during dynamical evolution. Their results may explain the reason why the value of α is small at a tail region. Figure 13 shows the schematic figure of the magnetic field configuration in SFO 22 suggested by comparisons between our numerical results and observations. The possible scenario is as follows. The magnetic field in SFO 22 was initially perpendicular to the UV radiation from the exciting star. Then, compression due to radiation-driven implosion made the magnetic field close to parallel to the UV radiation at the tail region.

Figure 13.

Figure 13. Schematic figures of magnetic field configuration in SFO 22 before undergoing compression by radiation-driven implosion (top) and after reaching the phase of quasi-stationary equilibrium (bottom).

Standard image High-resolution image

Very few attempts have been made to observationally study the magnetic fields of BRCs. Sridharan et al. (1996) performed optical polarimetry observations toward CG 22 and reported that the magnetic field is parallel to its tail. Their rough estimate gives a magnetic field strength of ∼30 μG. It is comparable to the magnetic field strength obtained by our numerical models (see Figure 10). Other optical polarimetry observations by Bhatt (1999) showed that the magnetic field in CG 30–31 is found to be nearly perpendicular to the cometary tails. Observational studies on the relations between the density structures and magnetic field configurations in BRCs are required to reveal the magnetic field effects on the evolution of BRCs.

6. CONCLUSIONS

Using the Nobeyama 45 m telescope, we observed BRC SFO 22 in the 12CO (J = 1–0), 13CO (J = 1–0), and C18O (J = 1–0) lines. Observed column density profiles were compared with those of numerical models for a photoevaporating cloud in quasi-stationary equilibrium state in order to investigate how the magnetic field and heating due to strong FUV radiation from the exciting star affect structures of BRCs. We summarize our main conclusions as follows:

  • 1.  
    From our radio observations, the column density profiles of SFO 22 along the line to its exciting star are nearly flat with the column density of ∼1022 cm−2.
  • 2.  
    Strong FUV radiation from the exciting star has little influence on the structure of SFO 22. Although enhanced thermal pressure due to strong FUV radiation slightly reduces the density of the cloud, its effects are not enough to reproduce the observed density structure of SFO 22.
  • 3.  
    The magnetic field strength and direction of the magnetic field strongly affect the structure of BRCs. The numerical model with an initial magnetic field strength of 5 μG shows the best agreement with the observations. When the magnetic field is nearly parallel to the UV radiation from the exciting star, the cloud has a steep column density profile at the head region and a flat column density profile at the tail region. When the magnetic field is nearly perpendicular to the UV radiation, the column density profile shows the opposite trend.

In this paper we only focus on the quasi-stationary equilibrium phase of the radiation-driven implosion model, and we will discuss the implosion phase in a subsequent paper. We also plan a further study using MHD simulation to establish a more realistic evolutionary model of BRCs.

The authors thank Kohji Sugitani, Shin-ya Nitta, and Hiroyuki Takahashi for helpful discussions. The authors thank the anonymous referee for constructive comments which improved the manuscript. We are grateful to the staff of Nobeyama Radio Observatory for their help during the observations. This work is supported by the Theoretical Institute for Advanced Research in Astrophysics (TIARA) operated under Academia Sinica in Taiwan. Numerical computations were in part carried out on the general-purpose PC farm at the Center for Computational Astrophysics, CfCA, of the National Astronomical Observatory of Japan. Numerical computations were carried out by using a workflow system, RENKEI-WFT, which is developed by the National Institutes of Informatics, Japan.

Please wait… references are loading.
10.1088/0004-637X/766/1/50