Magnetohydrostatic Equilibrium Structure and Mass of Polytropic Filamentary Cloud Threaded by Lateral Magnetic Field

and

Published 2021 April 22 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Raiga Kashiwagi and Kohji Tomisaka 2021 ApJ 911 106 DOI 10.3847/1538-4357/abea7a

Download Article PDF
DownloadArticle ePub

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

0004-637X/911/2/106

Abstract

Filamentary structures are recognized as a fundamental component of interstellar molecular clouds in observations made by the Herschel satellite. These filaments, especially massive filaments, often extend in a direction perpendicular to the interstellar magnetic field. Furthermore, the filaments sometimes have an apparently negative temperature gradient—that is, their temperatures decrease toward their centers. In this paper, we study the magnetohydrostatic equilibrium state of negative-indexed polytropic gas with the magnetic field running perpendicular to the axis of the filament. The model is controlled by four parameters: center-to-surface density ratio (ρc/ρs), plasma β of the surrounding gas, radius of the parent cloud ${R}_{0}^{{\prime} }$ normalized by the scale height, and the polytropic index N. The steepness of the temperature gradient is represented by N. We found that the envelope of the column density profile becomes shallow when the temperature gradient is large. This reconciles the inconsistency between the observed profiles and those expected from the isothermal models. We compared the maximum line mass (mass per unit length), above which there is no equilibrium, with that of the isothermal nonmagnetized filament. We obtained an empirical formula to express the maximum line mass of a magnetized polytropic filament as ${\lambda }_{{\rm{m}}{\rm{a}}{\rm{x}}}\simeq {\left[{\left({\lambda }_{0,{\rm{m}}{\rm{a}}{\rm{x}}}(N)/{M}_{\odot }{\rm{p}}{{\rm{c}}}^{-1}\right)}^{2}+{\left[5.9{\left(1.0+1.2/N\right)}^{1/2}\left({{\rm{\Phi }}}_{{\rm{c}}{\rm{l}}}/1\mu {\rm{G}}{\rm{p}}{\rm{c}}\right)\right]}^{2}\right]}^{1/2}$ M pc−1, where ${\lambda }_{0,\max }(N)$ represents the maximum line mass of the nonmagnetized filament and Φcl indicates one-half of the magnetic flux threading the filament per unit length. Although the negative-indexed polytrope makes the maximum line mass decrease, compared with that of the isothermal model, a magnetic field threading the filament increases the line mass.

Export citation and abstract BibTeX RIS

1. Introduction

Filamentary structures in molecular clouds have recently attracted much attention among researchers aiming to understand the earliest phase of star formation. The Herschel space observatory (Pilbratt et al. 2010) has revealed, by observing thermal dust emissions in the far-infrared and submillimeter ranges (André et al. 2010), that the filamentary structure is a basic component of nearby molecular clouds. Not only active star-forming regions, such as Aquila (Men'shchikov et al. 2010), Taurus (Palmeirim et al. 2013), and IC 5146 (Arzoumanian et al. 2011), but also inactive ones such as Polaris (Ward-Thompson et al. 2010) have indicated the presence of a filament system.

The magnetic field structure in molecular clouds can be studied in several ways, such as the near-infrared polarization of background stars and the polarization of thermal emissions from dust grains. Both methods are based on the fact that the dust grains are aligned along the magnetic field. In the former, background starlight is polarized parallel to the magnetic field. Many previous observations have indicated that the global magnetic field is nearly perpendicular to the main massive filaments (Chapman et al. 2011; Sugitani et al. 2011; Palmeirim et al. 2013; Kusune et al. 2016).

The far-infrared- and millimeter-wave polarization observations assume that the thermal emission from magnetically aligned dust is polarized in the direction perpendicular to the interstellar magnetic field. From the Planck all-sky survey, Planck Collaboration Int. XXXV (Planck Collaboration et al. 2016a) performed a statistical study for molecular clouds in the Gould belt, to determine whether the interstellar magnetic field is parallel or perpendicular to the major axis of filamentary structures. They found that the interstellar magnetic field is preferentially observed perpendicular to massive bright filaments, but filaments with low column density (striations) often extend in a direction parallel to the magnetic field. This trend is clearly seen in typical molecular clouds, such as Taurus, Lupus, and Chamaeleon–Musca.

Because the polarization is made by the magnetically aligned dust grains integrated along the line of sight, the apparent angle between the magnetic field and the filament is affected by their three-dimensional configuration (Tomisaka 2015; Planck Collaboration et al. 2016b; Doi et al. 2020; Reissl et al. 2021). For example, Doi et al. (2020) demonstrated that a major filament seems to extend parallel to the magnetic field observed in NGC 1333 by the James Clerk Maxwell Telescope, but the filament completely conforms to the commonly assumed configuration in which the magnetic field and filament are perpendicular to each other in three dimensions.

The stability of the interstellar filaments is often discussed using the mass per unit length, i.e., the line mass (Stodółkiewicz 1963; Ostriker 1964). In the case of an infinite cylindrical isothermal cloud with the central density ρc , the density profile ρ(r) is given analytically as

Equation (1)

where is the scale height, which is expressed as ${\ell }={c}_{s}/{\left(4\pi G{\rho }_{c}\right)}^{1/2}$ by using the isothermal sound speed cs and the gravitational constant G (Stodółkiewicz 1963; Ostriker 1964). Integration of Equation (1) along the radius r gives the line mass of a filament with the surface radius rs as

Equation (2)

The maximum line mass that can be supported against self-gravity is given by rs / as ${\lambda }_{\mathrm{iso},\mathrm{crit}}=2{c}_{s}^{2}/G$, which is called the critical line mass. The critical line mass is often used as a quantity that controls star formation inside the molecular cloud (Nagasawa 1987; Inutsuka & Miyama 1992; André et al. 2014). When the line mass λ exceeds the critical line mass λiso,crit in a filament, it contracts radially and begins star formation. For example, from the Herschel survey of the Aquila region, supercritical filaments with λ > λiso,crit contain most of the gravitationally bound prestellar cores (André et al. 2010).

Because star formation basically proceeds by gravitational contraction, knowing the conditions under which gravitational contraction begins leads to an understanding of the earliest phase of star formation. The equilibrium state of filaments has been studied from this standpoint. In addition to this, the magnetic field exists with the filamentary structures. The magnetic field also plays a central role in star formation; for a review, see, for example, Hennebelle & Inutsuka (2019). Thus, to understand the effect of this magnetic field on the equilibrium state, we should study the magnetohydrostatic equilibrium state.

Stodółkiewicz (1963) studied the equilibria of isothermal cylinders with the magnetic field parallel to their long axis. When the plasma β is constant, the critical line mass increases, owing to the increase of the scale length . In contrast, Fiege & Pudritz (2000) found that the toroidal magnetic field has an opposite effect to compress the filaments and reduces the critical line mass. However, these papers have discussed the case where the magnetic field is globally parallel to the filament.

Tomisaka (2014) studied the magnetohydrostatic equilibrium state of a filamentary isothermal cloud threaded by a lateral magnetic field. The study assumed a magnetized infinitely long cylindrical isothermal cloud and studied the effect of the magnetic field for the maximum line mass ${\lambda }_{\mathrm{iso},\max }$. Tomisaka (2014) numerically derived an empirical formula of the maximum line mass as

Equation (3)

where Φcl is one-half of the magnetic flux threading the filament per unit length. The study concluded that the maximum line mass supported against self-gravity is represented by the function of the magnetic flux Φcl, and when considering a filamentary cloud, it is necessary to account for the magnetic field.

To characterize the density of an axisymmetric filament, Plummer-like profiles are often used:

Equation (4)

where ρc is the central density, Rf is the core radius, and p is a density slope parameter (Nutter et al. 2008; Arzoumanian et al. 2011). The slope of the power-law distribution is determined by p, and the nonmagnetized isothermal cylinder corresponds to p = 4 (see Equation (1)). Density profiles observed by Herschel are reproduced well by a power-law distribution with an index around p ≃ 2. For example, this index is p ≃ 2.2 ± 0.4 for IC5146, p ≃ 2.4 ± 0.6 for Aquila, and p ≃ 2.3 ± 0.1 for Taurus (Arzoumanian et al. 2019).

Pineda et al. (2011), Hacar & Tafalla (2011), and Bourke et al. (2012) reported that the observed filament profile was fitted well with the isothermal model (Equation (1)) rather than p ≃ 2. However, because these are based on observations with high-density tracers (NH3, N2H+, and C18O), the obtained distribution may be affected by the abundance gradient. Even with the same Herschel data, Howard et al. (2019) pointed out that the shallow radial density gradient (p ≃ 2) seems to be affected by the smoothing and averaging inherent in its derivation. In addition, they claimed that, when analyzing each small local segment of the filament (of length 0.004 pc), the data indicate rather p = 4 than p = 2.

Although more deliberation may be needed, there currently is no strong evidence to reject p ≃ 2. Thus, in this paper, we explore the physical reason why the density profile of the filament is fitted with p ≃ 2. To explain a shallow density gradient like p ≃ 2, Toci & Galli (2015a) assumed a filament supported by thermal and nonthermal motions, then proposed the effect of the nonmagnetized gas obeying a nonisothermal equation of state.

Additionally, Arzoumanian et al. (2019) and Howard et al. (2019) reported that the temperature at the filament center is lower than that at the surface, i.e., the filament has a negative temperature gradient. This seems to be explained as the central part is shielded from the incoming interstellar radiation by the outer layer.

Based on this, Toci & Galli (2015a) studied the nonmagnetized infinite cylinder obeying the nonisothermal polytropic equation of state as follows:

Equation (5)

where γ, pg , K, and ρ are the polytropic exponent, gas pressure, proportional constant, and gas density, respectively. The polytropic exponent is often used as γ = 1 + 1/N, where N is the polytropic index. The polytropic index N represents how steep the temperature gradient is, and is negative (N ≤−1) when the filament has a negative temperature gradient. Toci & Galli (2015a) concluded that a negative polytropic index (− < N < −1) makes the density profile shallower than that of an isothermal model, N = − (see Figure 1 of their paper). Toci & Galli (2015b) also studied the effect of a helical magnetic field on the polytropic filament. They reported that the pitch angle, which is the angle between poloidal and toroidal magnetic fields, determines whether the magnetic field compresses or supports the polytropic filaments similarly to the isothermal ones (Fiege & Pudritz 2000).

Figure 1.

Figure 1. Model. Parent cloud (a) has uniform density with radius R0, which is threaded with the uniform magnetic field B0. Parent cloud is immersed in the external pressure pext and given a line mass of λ0. Starting from this state, we search the equilibrium state (b) while assuming that flux freezing preserves the initial mass distribution against the magnetic flux. Gas obeys the polytropic equation pg = K ρ1+1/N . The polytropic index N, the radius of the parent cloud ${R}_{0}^{{\prime} }\equiv {R}_{0}/\left[{c}_{\mathrm{ss}}/{\left(4\pi G{\rho }_{s}\right)}^{1/2}\right]$, the plasma beta ${\beta }_{0}\equiv {p}_{\mathrm{ext}}/({B}_{0}^{2}/8\pi )$, and the line mass $\lambda ^{\prime} \equiv {\lambda }_{0}/({c}_{\mathrm{ss}}^{2}/4\pi G)$ determine the equilibrium state. It should be noted that we use the center-to-surface density ratio ${\rho }_{c}^{{\prime} }\equiv {\rho }_{c}/{\rho }_{s}$ as the fourth parameter, instead of the line mass $\lambda ^{\prime} $, because it is easier to find the equilibrium state.

Standard image High-resolution image

In this paper, we present the numerical calculation for the equilibrium state of a magnetized filament that has a negative temperature gradient. The structure of this paper is as follows. In Section 2, we introduce the model and formulation for this calculation. We show the numerical results for this filament in Section 3. In Section 4, we discuss the effects of the magnetic field and the negative temperature gradient on the line mass and the filament structure. We summarize the results of this paper and provide conclusions in Section 5.

2. Method

The method to obtain the magnetohydrostatic structure for a polytropic gas is formulated based on the method for an isothermal gas (Tomisaka 2014).

2.1. Basic Equations

To derive the magnetohydrostatic configuration, we start from the following four equations. First, the polytropic equation is

Equation (6)

in which the meaning of the variables is the same as in Equation (5). The equation is based on the assumption that the pressure and the density are connected with the polytropic index. The second equation is a force balance equation between the Lorentz force, gravitational force, and pressure gradient, and is written as

Equation (7)

where c, j , B , and ψ represent the light speed, electric current density, magnetic flux density, and gravitational potential, respectively. The third equation is Poisson's equation for self-gravity, which is expressed as

Equation (8)

where G is the gravitational constant. The fourth equation is Ampère's law, which is written as

Equation (9)

We search for a solution for a filament extending infinitely in the z-direction and assume all the physical quantities depend on only (x, y). We introduce a magnetic flux function Φ, from which the magnetic flux density B = (Bx , By ) is given as

Equation (10a)

Equation (10b)

It should be noted that, in a 2D representation, the magnetic field line is given by a contour line of ${\rm{\Phi }}(x,y)=\mathrm{const}$. From Equation (9), the electric current j is rewritten as

Equation (11a)

Equation (11b)

Equation (11c)

where Δ2 is defined as

Equation (12)

Hereafter, ∇2 ≡ (∂/∂x, ∂/∂y) represents the two-dimensional differentiation operator. For the polytropic gas, the pressure term of Equation (7) is

Equation (13)

Using this equation, the force balance Equation (7) becomes

Equation (14a)

Equation (14b)

and when Bz = 0, these two equations reduce to

Equation (15a)

Equation (15b)

By taking the inner product of Equation (15) and B = (Bx , By ),

Equation (16)

is required in the direction parallel to the magnetic field lines. This means the quantity ψ + K(N + 1)ρ1/N is constant along a magnetic flux tube as

Equation (17)

where H(Φ) is the Bernoulli constant, which is a function dependent only on Φ. Along a magnetic tube given by a constant Φ, the density is calculated from the gravitational potential

Equation (18)

and then Equation (15) is rewritten as

Equation (19a)

Equation (19b)

Similarly to the method in Tomisaka (2014), we assume that forces are balanced inside the filament (Equation (19)). Outside the filament, we assume that a force-free magnetic field (the electric current j = 0) and hot tenuous gas (ρ = 0) exist. The extended gas confines the filament with its pressure (external pressure pext). Thus, the right-hand side of Equation (19) vanishes outside the filament. Finally, we can derive the two basic equations. Equation (19) leads to

Equation (20)

and with use of Equation (18), Poisson's Equation (8) is rewritten as

Equation (21)

We find the equilibrium state by solving these two second-order differential equations simultaneously via the self-consistent field method, but we need to know the value of H(Φ) at each magnetic field line.

2.2. Mass Loading

Here, we introduce a mechanism to derive H(Φ). In this paper, we assume that a large-scale magnetic field runs along the y-direction. We assume vertical symmetry at y = 0. Then, a line mass Δλ that is contained between two magnetic field lines, Φ and Φ + ΔΦ, is expressed as

Equation (22a)

Equation (22b)

Equation (22c)

where ys represents the y-coordinate of the filament surface and H(Φ) stays constant in the integration. This leads to the problem of finding an appropriate set of H(Φ) and ys (Φ) while at the same time satisfying the following equation:

Equation (23)

The left side of this equation is given as a model of the mass-to-flux ratio distribution, which is called mass loading.

We assume the central density of the filament as ρc and the potential as ψc . Then, Equation (17) gives the value of the Bernoulli constant for the central magnetic field line coinciding with the y-axis, which is specified by Φ = 0, as

Equation (24)

Equation (23) uses Equation (24) to obtain the mass loading on the central magnetic field line Φ = 0 as follows:

Equation (25)

where the upper boundary ys is given as a point where ψ = ψs . The surface potential is written as

Equation (26)

where the ρs is the density at the filament surface in the equilibrium state (see Figure 1(b)). Equation (25) indicates that, from a set of potentials Φ and ψ, the mass loading for the central magnetic field line ${\left.d\lambda /d{\rm{\Phi }}\right|}_{{\rm{\Phi }}=0}$ is obtained as a function of ρc .

In this paper, for the purpose of comparison with the isothermal model, we assume the following mass-loading distribution:

Equation (27)

which is the same assumed in Tomisaka (2014). In this equation, Φcl represents the magnetic flux threading the unit length of the filament and Φcl = R0 · B0, where R0 and B0 represent the initial radius of the filament and magnetic field strength, respectively, of the initial uniform magnetic field. This is realized when a uniform-density cylindrical filament is threaded with a uniform magnetic field, where −Φcl ≤ Φ ≤ Φcl represents the magnetic field line threading the filament, while Φ < −Φcl and Φ > Φcl represent the magnetic field lines not threading the filament (see Figure 1). In this paper, we refer to this filament with uniform density and uniform magnetic field as the "parent" cloud, which gives the mass loading in the filament in equilibrium. That is, we assume that the mass loading is determined by the parent cloud and is conserved by flux freezing.

For the magnetic field lines Φ ≠ 0, the equation becomes

Equation (28)

For a given Φ, ys (Φ) is chosen to satisfy the above equation, which is achieved with use of the bisection method of nonlinear equations.

2.3. Normalization Units

In this paper, physical variables are normalized with their quantities at the filament surface. We regard the following three quantities as fundamental: the external pressure pext, the surface density ρs , and the isothermal sound speed css at the filament surface. Then, for example, the scale length (L) is defined by the freefall time (tff) at the filament surface density (ρs ) and the isothermal sound speed (css) at the surface of the filament, as L = css tff. From the polytropic equation, the external pressure pext and the surface density ρs are related as

Equation (29)

The physical scales characterizing the system are given as in Table 1. We define the normalized variables as

Equation (30a)

Equation (30b)

Equation (30c)

Equation (30d)

Equation (30e)

Equation (30f)

Equation (30g)

Equation (30h)

where the prime represents the normalized variables. The density is normalized as

Equation (31)

Poisson's equation is rewritten as

Equation (32)

while Poisson's equation for magnetic flux function is given as

Equation (33)

The mass loading on the central magnetic field is given as

Equation (34)

and Equation (23) reduces to

Equation (35)

Table 1. Units used for Normalization

Unit of pressureExternal pressure, pext
Unit of densityDensity at the surface, ρs
Unit of timeFreefall time, ${t}_{\mathrm{ff}}={\left(4\pi G{\rho }_{s}\right)}^{-1/2}$
Unit of speed ${c}_{\mathrm{ss}}={\left({p}_{\mathrm{ext}}/{\rho }_{s}\right)}^{1/2}={\left(K{\rho }_{s}^{1/N}\right)}^{1/2}$
Unit of magnetic field strength ${B}_{u}={\left(8\pi {p}_{\mathrm{ext}}\right)}^{1/2}$
Unit of length $L={c}_{\mathrm{ss}}{t}_{\mathrm{ff}}={c}_{\mathrm{ss}}/{\left(4\pi G{\rho }_{s}\right)}^{1/2}={\left(K{\rho }_{s}^{-1+1/N}/4\pi G\right)}^{1/2}$
Unit of temperatureTemperature at the surface, Ts

Notes. Physical variables are normalized with their quantities at the filament surface. Because the freefall timescale (tff) and scale length (L) are defined on the filament surface, the unit of speed is taken as the isothermal sound speed (css) at the surface of the filament.

Download table as:  ASCIITypeset image

The two Poisson equations require boundary conditions. We impose the Dirichlet boundary condition on the outer numerical boundary given below. Far from the origin, we assume that the gravitational potential ψ converges to that realized for a line mass λ placed at the origin as

Equation (36)

The outer boundary condition for the magnetic potential is expressed as

Equation (37)

in which we assume that the magnetic field is connected to the uniform magnetic field with strength B0 far from the center. We then normalize the two potentials. The normalized value $\lambda ^{\prime} =\lambda /{\rho }_{s}{L}^{2}$ is used to reduce Equation (36) to

Equation (38)

The line mass λ is given as

Equation (39a)

Equation (39b)

Equation (39c)

If we know the ratio of mass to magnetic flux at the center, ${\left.d\lambda /d{\rm{\Phi }}\right|}_{{\rm{\Phi }}=0}$, we obtain the boundary value of $\psi ^{\prime} $ after calculating λ using Equations (38) and (39). The magnetic field potential is normalized as

Equation (40)

where β0 is a ratio of the external pressure pext to the magnetic pressure ${B}_{0}^{2}/8\pi $ and defined as

Equation (41)

2.4. Parameters

After the normalization, a solution is specified by four nondimensional parameters: ${{\rm{\Phi }}}_{\mathrm{cl}}^{{\prime} }$, β0, ${\rho }_{c}^{{\prime} }\equiv {\rho }_{c}/{\rho }_{s}$, and N (see Figure 1). The nondimensional magnetic flux ${{\rm{\Phi }}}_{\mathrm{cl}}^{{\prime} }$ is given as

Equation (42)

where ${R}_{0}^{{\prime} }$ is defined as the initial radius of uniform filament R0 normalized by the scale length L. Hereafter, we omit the prime, which indicates normalized quantities, unless the meaning is unclear.

2.5. Numerical Method

We solved Poisson's equation with the conjugate gradient method preconditioned with incomplete Cholesky factorization (ICCG). The number of grid points was chosen to be 641 × 641 or 1281 × 1281 and the grid spacing was chosen to be Δx = Δy = 0.1/16, 0.1/8, or 0.1/4. The outer numerical boundaries are placed at x = y = ±4 in the models with R0 = 1 and 2, and at x = y = ±8 in those with R0 = 5. We summarize the model parameters in Table 2.

Table 2. Model Parameters and Maximum Supported Line Mass

Model R0 β0 ${\rho }_{c\mathrm{Max}}$ Φcl ${\lambda }_{\max }$
    N = −3−5−10−100  N = −3−5−10−100
R1β111500103 103 20019.95713.4217.5423.90
R1β0.510.5500103 103 5001.4111.0614.4618.8425.60
R1β0.110.1500103 103 103 3.1616.3020.5125.3132.52
R1β0.0510.05500103 103 103 4.4719.83*25.2730.5438.03
R2β121103 103 103 100214.3617.7721.5726.84
R2β0.520.5103 † 103 103 5002.8317.2220.9424.9631.08
R2β0.120.1103 † 103 103 103 6.3229.9134.6339.8147.02
R2β0.0520.05103 † 103 103 103 8.9438.78*44.7550.8658.74
R5β151103 ◊ 103 • 103 • 103 • 527.7231.4835.3240.98
R5β0.550.5500 103 • 103 • 103 • 7.0735.9040.2244.6450.74
R5β0.150.1500 103 • 103 • 103 • 15.868.6176.1482.5991.37
R5β0.0550.05500 103 • 103 • 103 • 22.489.18*99.81*110.6120.5

Notes. The column ${\rho }_{c\mathrm{Max}}$ indicates that the solutions are obtained between ρc = 2 and ${\rho }_{c\mathrm{Max}}$. Numbers marked with the symbol * represent lower limits of ${\lambda }_{\max }$. The symbols †, ◊, and • represent respective grid spacings of 0.1/16, 0.1/8, and 0.1/4. In the models with † and ◊, the number of grid points is 1281 × 1281, and that of the model with • is 641 × 641. The rest of the models are calculated with grid points of 641 × 641 and a grid spacing of 0.1/8.

Download table as:  ASCIITypeset image

We verified our calculation by solving an approximate isothermal equilibrium state with the polytropic method and assuming N = −100 and −1000. From Equation (5), the polytropic indices N = −100 and N = −1000 correspond to the polytropic exponents γ (N = −100) = 0.99 and γ (N = −1000) = 0.999, both of which are close to the isothermal case of γ = 1. We compare the equilibrium state of the isothermal (Tomisaka 2014) and polytropic filaments (N = −100 and −1000) while paying attention to the line mass. In this comparison, the other parameters, i.e., the radius of the parent cloud R0 = 2 and the plasma beta β0 = 0.1, are constant. When the central density is ρc = 103, the line mass of the polytropic filament is λN=−100 = 47.02 and λN=−1000 = 47.96, while the isothermal one is λiso = 48.07. When the central density is ρc = 102, the corresponding line masses are λN=−100 = 43.31, λN=−1000 = 44.00, and λiso = 44.08.

The line mass of polytropic filaments is slightly lower than the isothermal one. However, it is clearly shown that the line mass converges to the isothermal value when N moves to −. Thus, our calculation reproduces a line mass close to the isothermal one when the polytropic index N → −. Hereafter, we assume the results with N = −100 to be the isothermal model.

3. Results

3.1. Comparison of Polytropic (N = –3) and Isothermal (N = –100) Filaments

In this section, we compare the density profile and the line mass of the polytropic (N = −3) and isothermal (N = −100) filaments. In the comparison, other parameters of these filaments, the radius of the parent cloud R0 = 1 and the plasma beta β0 = 0.1 are constant.

First, we begin with the density distribution of the equilibrium state. We show the cross sections of the polytropic filaments in Figure 2(a)–(c) and the isothermal filaments in panels (d)–(f). In this paper, we refer to the x- and y-coordinates of the cloud surface crossing the x- and y-axes as the "width" xs and "height" ys of the filament, respectively. It is shown that both of these filaments become flatter as the central density increases: the height of the filament shrinks, while the width remains almost unchanged. This is due to the character of the Lorentz force, which works in the perpendicular direction to the magnetic field line but does not work in the parallel direction. Because extra force to support the filament is working in the direction perpendicular to the magnetic field, the isodensity contours of the cross section shrink mainly in the y-direction and appear flat.

Figure 2.

Figure 2. Comparison of polytropic (panels (a)–(c): N = −3) and isothermal (panels (d)–(f): N = −100) filaments. This figure shows the cross section of the equilibrium state on the xy plane. Other parameters are constant: radius of the parent cloud R0 = 1 and plasma beta β0 = 0.1. We show three models with different central densities ρc as ρc = 10 ((a) and (d)), ρc = 100 ((b) and (e)), and ρc = 500 ((c) and (f)). Solid lines show the isodensity contours, and each contour level is chosen to be ρ = 2, 3, 5, 10, 20, 30, 50, 100, 200, 300, 500, and 1000, respectively, from outside to inside. Dotted vertical lines are magnetic field lines. The line masses of these filaments are λN=−3 = 10.36 (a), 15.10(b), 16.30(c) and λN=−100 = 17.90 (d), 29.47(e), 32.18(f).

Standard image High-resolution image

Figure 2 shows that the cross section of the polytropic filament is flatter than the isothermal one when two with the same central density are compared. However, the outer part's gas scale height in the y-direction of the polytropic filament is nearly equal to that of the isothermal one. For example, panel (b) shows that the polytropic filament has a height of ys ≃ 0.39 on the symmetric y-axis. However, the surface inflates outward, and the height of the surface reaches ≃0.60 near x ≃ 0.80. Thus, this polytropic filament has a maximum height of ∼0.60. In contrast, the corresponding isothermal model (panel (e)) does not show such inflation (ys ≃ 0.64 and maximum height ≃0.66). As is shown, the maximum height of the polytropic filament is nearly the same as that of the isothermal filament. This can be understood by accounting for the temperature near the surface of the polytropic filament, which is not very different from the temperature of the isothermal filament, although the polytropic filament has a lower central temperature compared with the isothermal filament.

Figure 3 shows the density profiles on the x- and y-axes. In particular, the density profile on the y-axis clearly shows the effect of different N values. Figure 3 shows that the density distribution is divided into two parts: an inner core with an almost constant density and an outer envelope in which the density decreases with increasing distance from the center. This figure shows that both the isothermal and the polytropic filaments have power-law envelopes, except for the envelopes cutting along the x-axis for the ρc = 10 models. In terms of the distance to the surface from the center, the polytropic filament is more compact than the isothermal one in the y-direction. In addition, the ρ(y) distribution indicates that the density slope is shallower than the isothermal slope. Although these two results seem to be inconsistent, this is natural if we consider that the compactness of a polytropic filament comes from the fact that it has a smaller core than an isothermal filament.

Figure 3.

Figure 3. Comparison of the density profiles on the x- and y-axes for polytropic (N = −3) and isothermal (N = −100) filaments. Other parameters are constant: R0 = 1 and β0 = 0.1. Vertical axis shows the density, and the horizontal axis shows the distance from the center. Red and cyan curves represent the models with N = −3 and N = −100, respectively. Solid and dashed curves correspond to the density profiles on the y- and x-axes, respectively. These panels correspond to different central densities: ρc = 10 (a), ρc = 100 (b), and ρc = 500 (c).

Standard image High-resolution image

In contrast, density profiles on the x-axis are almost identical. For example, the height of the polytropic filament on the y-axis is 40% smaller than the isothermal height, while the x-axis width is 20% wider than the isothermal width, when we compare density distributions with the same central density ρc = 500.

Next, we pay attention to the difference between the line mass of each filament. For the central densities of ρc = 10, 100, and 500, the line masses of the polytropic and the isothermal filaments are obtained to be λN=−3 = 10.36, 15.10, 16.30, and λN=−100 = 17.90, 29.47, 32.18, respectively. In both the polytropic and isothermal models, the line mass increases as the central density increases. Meanwhile, comparison of filaments with the same central density shows that the polytropic filament is less massive than the isothermal one. This is explained by the fact that the central pressure, which supports the filament against self-gravity, of the polytropic filament is smaller than that of the isothermal one. For example, comparing two models with ρc = 100, the central pressure of the polytropic model is only pc = 1002/3, which is only ∼21.5% of the central pressure of the isothermal model. Note that these properties come from the nature of the negative-indexed polytropic gas that is immersed in the same ambient gas pressure.

3.2. Comparison of the Radius of the Parent Cloud

In this section, we address the effect of the radius of the parent cloud R0, which controls the magnetic flux threading the filament. The other parameters are constant at N = −3 and β0 = 0.1.

Figure 4 shows the cross sections of the models of R0 = 2 [(a) − (c)] and R0 = 5 [(d) − (f)] for respective central densities ρc = 10, 100, and 500 (the model with R0 = 1 is shown in the upper row of Figure 2). When ρc = 500, the height ys of the filament on the y-axis is equal to ys = 0.238 (R0 = 1), ys = 0.213(R0 = 2), and ys = 0.188(R0 = 5) for the three different R0 values, respectively. In contrast, the half-width xs on the x-axis is equal to xs = 0.888 (R0 = 1), xs = 1.75(R0 = 2), and xs = 4.41(R0 = 5), respectively. Thus, the aspect ratio xs /ys for R0 = 1, 2, and 5 is equal to xs /ys = 3.73, 8.22, and 23.5, respectively. Thus, the aspect ratio is an increasing function of R0.

Figure 4.

Figure 4. Comparison of the models with different R0 values: R0 = 2 [(a)–(c)] and R0 = 5 [(d)–(f)]. Other parameters are constant: N = −3 and β0 = 0.1. The central density is equal to ρc = 10 [(a) and (d)], ρc = 100 [(b) and (e)], and ρc = 500 [(c) and (f)]. Solid contour lines indicate that the isodensity is similar to that in Figure 2. Dotted vertical lines are the magnetic field lines. The line masses of these filaments are ${\lambda }_{{R}_{0}=2}=17.51\,({\rm{a}}),\,26.68\,({\rm{b}}),\,29.42\,({\rm{c}})$, and ${\lambda }_{{R}_{0}=5}=39.31\,({\rm{d}}),\,61.50\,({\rm{e}}),\,68.61\,({\rm{f}})$.

Standard image High-resolution image

For the nonmagnetic model of β0 = , we expected the cross section to be round. However, the shape of the cross section is flat in the above models. The magnetic field supports the filament in the x-direction, but does not play a role in the y-direction. The average ratio of the Lorentz force to the thermal pressure force is equal to 3.27 (R0 = 1), 6.25(R0 = 2), and 15.3(R0 = 5) for ρc = 500, respectively, measured on the x-axis. Thus, the Lorentz force is stronger than the thermal pressure, especially for the model with R0 = 5. In addition, comparing three models with ρc = 10, we found an aspect ratio xs /ys = 1.25 (R0 = 1), 2.889(R0 = 2), and 8.083(R0 = 5) for the three different R0 values, respectively. Models of ρc = 100 indicate an aspect ratio xs /ys = 2.350 (R0 = 1), 5.407(R0 = 2), and 15.167(R0 = 5), respectively. This shows that the aspect ratio increases as the central density increases when R0 is the same. Figure 5 shows the density profiles on the x- and y-axes. The density profile on the y-axis is more compact than that on the x-axis, and the filament is flat. Comparison of the models with the same central density shows that the slope of the density profile on the y-axis is almost the same for the three different R0 values. In contrast, the density profiles on the x-axis are not the same. The core radius on the x-axis increases as R0 increases, and as a result, the distance to the surface also increases with increasing R0.

Figure 5.

Figure 5. Same as Figure 3, but for comparison of models with different R0 values. Line color represents R0 as R0 = 1 (blue), R0 = 2 (green), and R0 = 5 (red). Other parameters are constant: N = −3 and β0 = 0.1. Panels (a), (b), and (c) correspond to the models with ρc = 10, 100, and 500, respectively. Dashed and solid curves correspond to the density profile on the x- and y-axes, respectively.

Standard image High-resolution image

Next, we examine the difference in the line mass. Figure 6 shows the relation between the line mass and the central density for various N and R0 values with constant plasma beta β0 = 0.1. Comparison of models with the same central density and polytropic index shows that the line mass increases as R0 increases. This suggests that the supported line mass is controlled by the magnetic flux Φcl = R0 · B0. This property is also valid for other polytropic indices. Figure 6 also shows that the line mass decreases with increasing polytropic index from N = −100 to −3, which is also discussed in Section 3.1.

Figure 6.

Figure 6. Line mass plotted against the central density. Plasma beta is constant at β0 = 0.1. Vertical and horizontal axes represent the line mass and the central density, respectively. The colors blue, green, and red correspond to the models with R0 = 1, 2, and 5, respectively. The symbols ◦, □, ⋆, and △ correspond to the models with N = −100, −10, −5, and −3, respectively.

Standard image High-resolution image

3.3. Effect of Plasma Beta

Next, we compare the models with different β0 values. Other parameters are fixed: N = −3, R0 = 2, and ρc = 100.

Figure 7 shows the cross section of the equilibrium state. Each panel corresponds to a different β0 value: (a) β0 = 1, (b) β0 = 0.5, (c) β0 = 0.1, and (d) β0 = 0.05. When β0 is small and the magnetic field is strong, the magnetic field line retains its initial shape. Because the filament is sufficiently supported by the strong magnetic field, the surface of the filament is close to R0 on the x-axis (xs R0).

Figure 7.

Figure 7. Comparison of the solution models with different β0 values. Other parameters are constant: N = −3, R0 = 2, and ρc = 100. Panels (a), (b), (c), and (d) correspond to β0 = 1, 0.5, 0.1, and 0.05, respectively. The line masses of the respective models are ${\lambda }_{{\beta }_{0}=1}=14.26$ (a), ${\lambda }_{{\beta }_{0}=0.5}=17.22$ (b), ${\lambda }_{{\beta }_{0}=0.1}=26.68$ (c), and ${\lambda }_{{\beta }_{0}=0.05}=31.16$ (d).

Standard image High-resolution image

Figure 8 shows the density profiles on the x- and y-axes for the same models shown in Figure 7. The density profile on the y-axis is slightly affected by β0. In contrast, the density profile on the x-axis becomes steep in the models with low β0. The strong Lorentz force extends the core radius, but the width xs is not strongly affected by β0. Thus, the thickness of the envelope shrinks, which makes the slope steep.

Figure 8.

Figure 8. Density profiles on the x- and y-axes for the models shown in Figure 7. Other parameters are constant: N = −3, R0 = 2, and ρc = 100. Line colors represent the models with different β0 values as β0 = 1 (red), β0 = 0.5 (magenta), β0 = 0.1 (green), and β0 = 0.05 (blue). Solid and dashed curves show the density profiles along the y- and x-axes, respectively.

Standard image High-resolution image

Figure 9 shows the relation of the line mass and the central density for filaments with N = −3 (△) and N = −100 (◦), in which different line colors represent different β0 values. All the models have the same R0 = 2. Comparison of magnetized and nonmagnetized polytropic filaments (△for N = −3) indicates that the line mass of the magnetized filament is heavier than the nonmagnetized one (black symbols and solid curve). Results for the isothermal filaments (N = −100) are the same. The line mass of the polytropic filament with N = −3 (△) is smaller than that with N = −100 (◦) when models with the same β0 and ρc are compared. This reflects the fact that the line mass of the negative-indexed polytropic filament is less massive than that for the isothermal filament. However, when the magnetic field is strong, the line mass of the magnetized N = −3 filament is even larger than that of the nonmagnetized isothermal filament (N = −100: gray symbols and solid curve).

Figure 9.

Figure 9. Relation between the line mass and the central density for polytropic (N = −3) and isothermal (N = −100) filaments. Line colors represent models with different β0 values: β0 = 1 (red), 0.5 (magenta), 0.1 (green), and 0.05 (blue). Thin colored curves with ◦symbols and thick colored curves with △ symbols represent, respectively, N = −100 and N = −3 models. The black (N = −3) and gray (N = −100) solid curves show results obtained from the nonmagnetized Lane–Emden equation.

Standard image High-resolution image

4. Discussion

4.1. Maximum Line Mass

Section 3.1 shows that the line mass decreases as N increases from −100 to −3. In Section 3.2, it is shown that the line mass increases with increasing R0. The line mass increases with decreasing β0, as shown in Section 3.3. Thus, we expect the line mass to be determined by the magnetic flux ΦclR0 · B0 and N. Here, we discuss how the maximum line mass ${\lambda }_{\max }$ is expressed by the magnetic flux Φcl.

The maximum line mass ${\lambda }_{\max }$ represents the maximum allowable line mass of a filament that is in equilibrium. When the line mass λ exceeds ${\lambda }_{\max }$, there is no equilibrium state and we call the filament "supercritical." In contrast, a filament with $\lambda \lt {\lambda }_{\max }$ is called "subcritical," and its solution is discussed in the previous section. For a review, see André et al. (2014).

Although from its definition, ${\lambda }_{\max }$ is calculated as the slope ${\left(\partial \mathrm{log}\lambda /\partial \mathrm{log}{\rho }_{c}\right)}_{N,{R}_{0},{\beta }_{0}}=0$, considering numerical errors, we regard ${\lambda }_{\max }$ to be achieved when ${\left(\partial \mathrm{log}\lambda /\partial \mathrm{log}{\rho }_{c}\right)}_{N,{R}_{0},{\beta }_{0}}\lt 0.05$ is satisfied. 3

Figure 10 plots ${\lambda }_{\max }$ against Φcl for various N values. This shows that ${\lambda }_{\max }$ increases with Φcl, and the slope seems almost the same at large Φcl ≳ 10. We approximated the curves by using a function ${\lambda }_{\max }=\sqrt{{\lambda }_{0,\max }^{2}+A{{\rm{\Phi }}}_{\mathrm{cl}}^{2}}$, where ${\lambda }_{0,\max }$ corresponds to the maximum line mass of a nonmagnetized polytropic filament, and A represents the slope, which is determined by fitting. The value of A for various N values is obtained as A = 32.0 (N = −100), 26.4(N = −10), 23.5(N = −5), and 18.9(N = −3). From these four points, A is fitted as A = 31.5 + 39.0/N, and the accuracy of this fitting is within 5%. Thus, the normalized line mass is expressed with ${\lambda }_{0,\max }^{{\prime} }$ and ${{\rm{\Phi }}}_{\mathrm{cl}}^{{\prime} }$ as

Equation (43)

where ${\lambda }_{0,\max }^{{\prime} }$ for various N values is ${\lambda }_{0,\max }^{{\prime} }(N=-100)\,=23.2$, ${\lambda }_{0,\max }^{{\prime} }(N=-10)=17.5$, ${\lambda }_{0,\max }^{{\prime} }(N=-5)=14.4$, and ${\lambda }_{0,\max }^{{\prime} }(N=-3)=12.0$. Because the line mass and magnetic flux are normalized by ${c}_{\mathrm{ss}}^{2}/4\pi G$ and ${c}_{\mathrm{ss}}^{2}/{\left(G/2\right)}^{1/2}$, as in Table 1, we obtain a dimensional form of Equation (43) as

Equation (44)

Although the critical mass-to-flux ratio of the magnetized isothermal filamentary cloud is ${\left({G}^{1/2}\lambda /{{\rm{\Phi }}}_{\mathrm{cl}}\right)}_{\mathrm{crit}}=0.24$ (Tomisaka 2014), the value for the negative-indexed polytropic filament (N = −3) is approximately equal to 0.24. Finally, Equation (44) is rewritten as

Equation (45)

where the maximum line mass of a nonmagnetized polytropic filament at each N value is obtained as

Equation (46)

Thus, we derived an empirical formula of the maximum line mass for magnetized polytropic filaments as Equation (45). For example, an N = −3 filament shows that the magnetic contribution for the line mass becomes dominant when ${{\rm{\Phi }}}_{\mathrm{cl}}\gt 4.6\,\mathrm{pc}\,\mu {\rm{G}}{\left({c}_{\mathrm{ss}}/190{\rm{m}}{{\rm{s}}}^{-1}\right)}^{2}$.

Figure 10.

Figure 10. Maximum line mass plotted against the magnetic flux. Color represents different polytropic indices N: cyan (N = −100), blue (N = −10), green (N = −5), and red (N = −3). The × symbols mean that the value is reliable, i.e., obtained under the condition of $0\leqslant \partial \mathrm{log}\lambda /\partial \mathrm{log}{\rho }_{c}\leqslant 0.05$, while ◊ symbols indicate that the value is the lower limit, $\partial \mathrm{log}\lambda /\partial \mathrm{log}{\rho }_{c}\gt 0.05$. Filled circles (•) represent the maximum line mass of the nonmagnetized polytropic filament Φcl = 0.

Standard image High-resolution image

4.2. Column Density Distribution

Figures 3, 5, and 8 show the density profiles of models with various ρc , N, R0, and β0 values. The density profile on the y-axis is almost the same for three different R0 and four different β0 values. The slope of the profile on the y-axis becomes shallow when we increase N from −100 to −3. Thus, the slope of the density profile is controlled only by the polytropic index N in the y-direction, where the Lorentz force does not work.

This can be understood as follows. In the y-direction, because the Lorentz force does not play a role, the density distribution is governed by the pressure distribution (and self-gravity). A negative temperature gradient from the surface to the center has the effect of extending the envelope, and the temperature gradient increases from an isothermal equation of state (N = −100) to a polytropic one (N = −3).

Conversely, the slope on the x-axis becomes shallow only for the models with a weak magnetic field (Figure 8), while the slope slightly changes with R0 (Figure 5).

We pay attention to these characteristics when considering how to reproduce the observed column density profile. To characterize the density of the axisymmetric filament, Plummer-like profiles are often used, such as Equation (4). Accordingly, the observed column density distribution is fitted with the function

Equation (47)

which is also a Plummer-like function, where σ0, Rf , and p—the central column density, the core radius, and the density slope parameter, respectively—are three fitting parameters (Nutter et al. 2008; Arzoumanian et al. 2011). We obtain the column density by integrating the numerical solution of the density distribution as

Equation (48a)

Equation (48b)

where σ(x) and σ(y) represent the column densities observed from the parallel and perpendicular directions, respectively, with respect to the magnetic field.

As shown in Equation (1), the column density profile of an isothermal filament in hydrostatic equilibrium follows p = 4 (Stodółkiewicz 1963). However, as is summarized in Section 1, Herschel observations indicate that almost all the filaments follow p ≃ 2. Although some researchers have argued that the cylindrical dynamical contraction explains the observed shallow column density slope of p ≃ 2 (e.g., Kawachi & Hanawa 1998), in the present paper, we investigate whether a hydrostatic filament having a negative temperature gradient forms the observed shallow column density slope.

In Figure 11, we show the column density integrated along the direction of the y- and x-axes for various N and β0 values. Other parameters are constant at R0 = 2 and ρc = 100. We determined three parameters in the Plummer-like function (the slope index p, the column density at the center σ0, and the core radius Rf in Equation (47)) via the least-squares method. The least squares are calculated only for the region of σσ0/10. This restriction comes from accounting for the dynamic range of the observed column density above the fore- and background column density (Arzoumanian et al. 2019).

Figure 11.

Figure 11. Fitted Plummer-like distributions: Each panel represents the column density distributions σ(x) [(a) and (b)] and σ(y) [(c) and (d)] for various N values. Panels (a) and (c) correspond to β0 = 1, while (b) and (d) correspond to β0 = 0.05. Other parameters are constant: R0 = 2 and ρc = 100. Dashed curves represent the column density distribution of magnetohydrostatic filaments, which are integrated along the y- and x-directions. Fitted Plummer-like distributions are shown by thin colored solid curves. Color represents different polytropic indices N: cyan (N = −100), blue (N = −10), green (N = −5), and red (N = −3).

Standard image High-resolution image

Figure 11(a) and (b) presents the column density profiles of σ(x) that correspond to the profile in the direction in which the Lorentz force is effective. When β0 = 1, p reaches 2 as N goes from −100 to −3, p(N = −100) = 4.86 → p(N = −3) = 2.48 (panel a). Conversely, p does not show such convergence when β0 = 0.05. For example, the model with N = −3 (red curve of Figure 11(b)) indicates that the range of the power-law column density distribution is very narrow, and just outside of this, a sharp density decrement is observed. If we try to fit this column density distribution with a Plummer-like function, this gives an artificially large power-law index p.

In conclusion, the slope of the column density profile becomes shallow due to the temperature gradient for a model with a weak magnetic field. In contrast, we found that a strong magnetic field makes Rf large and worsens the fitting with the Plummer-like function (47).

Next, Figure 11(c) and (d) corresponds to σ(y), which indicates the column density distribution in the direction in which the Lorentz force does not play a role. For both β0 = 1 (panel c) and β0 = 0.05 (panel d), the slope index p reaches 2 as N changes from −100 to −3. This resembles the relation obtained for the nonmagnetized polytropic filament studied by Toci & Galli (2015a). In the direction in which the Lorentz force is less important, the negative temperature gradient toward the center plays a role in making the density slope shallow, even in a magnetized filament.

4.3. Column Density Distribution Depending on the Line of Sight

In Figure 11, we plot the column density distributions observed from the y-direction, σ(x), and that from the x-direction, σ(y). For comparison with observations, in this section we study the dependence of power-law index p and core radius Rf on the line of sight direction θ, which is defined as the angle between the line of sight and the x-axis. Defining the angle between the line of sight and the x-axis as θ, we rotated the density distribution at −θ. With the rotated density distribution integrated along the x-axis, σx (y) ≡ ∫ρ dx gives the column density distribution observed from this line of sight. In this case, the column density profiles σ(y) and σ(x) are obtained when θ = 0° and θ = 90°, respectively. Fitting the rotated σx (y) with the Plummer-like function of Equation (47), we obtain the power-law index p and the core size Rf depending on θ, as shown in Figure 12. The parameters of the model are R0 = 2, N = −3, β0 = 1, and ρc = 100.

Figure 12.

Figure 12. Core radius Rf and density slope parameter p plotted against θ, which is the angle between the line of sight and the x-axis. Left and right vertical axes show p and Rf , respectively. Solid and dashed curves show the values of Rf and p at each θ, respectively.

Standard image High-resolution image

Figure 12 shows that the power-law index p (dashed curve) is restricted to a narrow range of 2.48 ≲ p ≲ 2.73. Thus, if we measure p, the line-of-sight direction is hardly determined. In other words, this seems to explain the reason why filaments commonly have p ≃ 2, even if the line of sight and thus the angle θ must be chosen randomly for each observational target. In contrast, the core size Rf (solid curve) changes smoothly from 0.07 (θ = 0°) to 0.26 (θ = 90°). Because Rf is strongly dependent on θ, we can distinguish whether the line of sight is nearly perpendicular (θ ≃ 0°) or parallel (θ ≃ 90°) to the magnetic field.

We now propose a way to distinguish whether θ = 0° or θ = 90° when the line of sight is perpendicular to the filament long axis. In the nonmagnetized and thus symmetric filament, Equation (47) indicates that the central column density is equal to σ0 = S ρc Rf , where the numerical factor $S\,=2{\int }_{0}^{\infty }{\left(1+{\zeta }^{2}\right)}^{-p/2}d\zeta $ equals S = π for p = 2 and S = π/2 for p = 4. This means that, in the axisymmetric model, the central column density is given as the central density times the scale length of the column density in the direction perpendicular to the line of sight.

In the nonaxisymmetric configuration, we define the effective central density as

Equation (49)

Table 3 shows the quantity calculated assuming S = π for the models shown in Figure 11. Because all the models have the same central density ρc = 100, ${\rho }_{c}^{\mathrm{eff}}$ derived from σ(x) is smaller than the true ρc . For N = −3, ${\rho }_{c}^{\mathrm{eff}}$ derived from σ(y) is much larger than the true ρc , but ${\rho }_{c}^{\mathrm{eff}}$ derived from σ(x) is much smaller than that. When the central density is observationally estimated, for example, by using the critical density of the molecular line transitions, we can compare the ρc and ${\rho }_{c}^{\mathrm{eff}}$ estimated from the central column density σ0 and the column density scale length Rf . For a filament with N = −3, ${\rho }_{c}^{\mathrm{eff}}\gg {\rho }_{c}$ for σ(y). Therefore, when we observe ${\rho }_{c}^{\mathrm{eff}}\gg {\rho }_{c}$, this indicates that the line of sight is perpendicular to the magnetic field. Conversely, ${\rho }_{c}^{\mathrm{eff}}\ll {\rho }_{c}$ indicates that the line of sight is parallel to the magnetic field. Based on this, when we observe the central density, central column density, and core radius of the filament, we can evaluate the angle between the line of sight and the magnetic field line.

Table 3. Effective Central Density ${\rho }_{c}^{\mathrm{eff}}$

β0 N σ(y) σ(x)
   Rf σ0 ${\rho }_{c}^{\mathrm{eff}}$ Rf σ0 ${\rho }_{c}^{\mathrm{eff}}$
1−1000.265567.40.463725.6
1−100.18631110.382924.3
1−50.12621650.312222.6
1−30.07602730.251519.1
0.05−1000.241411870.923110.7
0.05−100.161452891.01257.88
0.05−50.111514371.32194.58
0.05−30.071627373.26131.27

Notes. Here, σ(x) and σ(y) represent the column density distributions obtained by integrating the density in the directions parallel and perpendicular to the magnetic field, respectively.

Download table as:  ASCIITypeset image

5. Summary and Conclusions

We used the negative-indexed polytropic model to investigate the magnetohydrostatic equilibrium state of an interstellar filament with a lateral magnetic field and negative temperature gradient. Our findings are as follows:

  • 1.  
    Increasing the polytropic index from N = −100 to −3 flattens the filament cross section in a direction parallel to the magnetic field. When the density profiles of polytropic and isothermal filaments are compared in a direction parallel to the magnetic field, the envelope of polytropic filament is shallower. The line mass of a polytropic filament is less massive in comparison to that of an isothermal filament when the filaments have the same central density and surface temperature.
  • 2.  
    When the radius of the parent cloud increases from R0 = 1–5, the aspect ratio of the cross section (major-to-minor axis ratio) also increases. Comparison of models with the same central density shows that the slope of the density profile parallel to the magnetic field is almost the same for three different R0 values. In contrast, the density profiles perpendicular to the magnetic field are not the same, because the core radius in that direction increases when R0 increases. The line mass increases with R0 when we compare models with the same central density.
  • 3.  
    Over the whole range of β0 = 0.05 − 1, we found that the width of the filament in the direction perpendicular to the magnetic field is almost the same as that at R0. In this direction, a model with stronger magnetic field has a larger core radius than that of a weak magnetic model. Thus, in such a model, the density profile in the direction perpendicular to the magnetic field has a steep slope outside the core. Meanwhile, the density profile in the direction parallel to the magnetic field is almost the same irrespective of β0. The line mass becomes heavy with a small β0 (strong magnetic field).
  • 4.  
    We found that the maximum line mass increases with the magnetic flux, and obtained the ratio of critical mass to magnetic flux: $\sim {\left(0.10+0.12/N\right)}^{1/2}{G}^{-1/2}$.
  • 5.  
    We conclude that a shallower column density profile is produced by a negative temperature profile in a magnetized filament. We succeeded in reproducing the observed column density profiles—especially in the direction where the Lorentz force is not effective, or in the model with a weak magnetic field. In the direction where the Lorentz force is effective, this mechanism does not work for a model with a strong magnetic field.
  • 6.  
    We have proposed a way to estimate the angle between the line of sight and the magnetic field line. We found that the core radius Rf is strongly dependent on this angle. This relation may help us distinguish whether the line of sight is nearly perpendicular or parallel to the magnetic field.

The authors would like to thank Dr. K. Iwasaki for discussions on the model formulation and for careful reading of the manuscript. This work was supported in part by a Grant-in-Aid for Scientific Research (C) (No. 19K03919) from the Japan Society for the Promotion of Science (JSPS), in 2019–2021.

Footnotes

  • 3  

    As indicated in Figure 9, although in most of the models, λ(ρc ) monotonically increases with increasing ρc , some models [R2β1 (N = −3) and R2β0.5 (N = −3)] have apparent peaks. These models with peaks enable us to estimate the error in ${\lambda }_{\max }$ for a model in which λ does not have a peak over the whole range of calculation, that is, $2\leqslant {\rho }_{c}\leqslant {\rho }_{c,\max }$. From model R2β1 (N = −3), the criteria ${\left(\partial \mathrm{log}\lambda /\partial \mathrm{log}{\rho }_{c}\right)}_{N,{R}_{0},{\beta }_{0}}=0.05$ and = 0.1 yield line masses approximately 1% and a few percent smaller than the true maximum line mass, respectively.

Please wait… references are loading.
10.3847/1538-4357/abea7a