This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

Alignment of Irregular Grains by Radiative Torques: Efficiency Study

, , and

Published 2021 May 25 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Joonas Herranen et al 2021 ApJ 913 63 DOI 10.3847/1538-4357/abf096

Download Article PDF
DownloadArticle ePub

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

0004-637X/913/1/63

Abstract

We study the efficiency of grain alignment by radiative torques (RATs) for an ensemble of irregular grains. The grains are modeled as ensembles of oblate and prolate spheroids, deformed as Gaussian random ellipsoids, and their scattering interactions are solved using numerically exact methods. We define the fraction of the grains that both rotate fast and demonstrate perfect alignment with grain long axes perpendicular to the magnetic field. We quantify a factor related to the efficacy of alignment and show that it is related to a ${q}_{\max }$ factor of the analytical model of the RAT theory. For the interstellar radiation field, our results indicate that the degree of RAT alignment can reach ∼0.5, which may be sufficient to explain observations even if grains do not have magnetic inclusions.

Export citation and abstract BibTeX RIS

1. Introduction

Measuring the polarization from aligned grains is the most common way of studying magnetic fields in galactic disks and molecular clouds (see Crutcher et al. 2010). Polarization arising from aligned dust interferes with attempts to measure enigmatic B-modes of cosmological origin (Philcox et al. 2018).

Shortly after the first aligned grains were detected by Hiltner (1949) and Hall (1949), the first theories of alignment were proposed by Davis & Greenstein (1951) for paramagnetic alignment and Gold (1952) for mechanical alignment. These became the two mainstream directions of the research for several decades, with paramagnetic alignment being the clear favorite (see Lazarian 2003 for a historic review). Significant contributions to understanding of the complex dynamics of grains in the process of alignment were made by L. Spitzer (Jones et al. 1967; Spitzer & McGlynn 1979) and E. Purcell (Purcell 1969, 1979). The grains in the studies were assumed to be regular, i.e., spherical or ellipsoidal.

A very different approach was proposed in Dolginov & Mytrophanov (1976a) and Dolginov & Mytrophanov (1976b), who suggested that an essential piece of physics related to alignment of reallistically irregular grains was missing in the theory. The authors proposed that irregular grains can demonstrate helicity in the process of their interaction with radiation, and the interaction can induce their alignment. However, these studies were not able to demonstrate the efficiency of their mechanism, and as was shown by later research (Hoang & Lazarian 2009), the toy model described by the authors was unable to exhibit alignment by radiation. In addition, the theoretical prediction that the direction of grain alignment, i.e., that the grains become aligned parallel or perpendicular to the magnetic field depending on the angle between radiation and the magnetic field, did not agree with polarization measurements (Andersson et al. 2015). As a result, these pioneering studies were mostly ignored for nearly two decades.

It became possible to test the idea of the irregular grain interaction with radiation due to the progress in computations as well as due to the development of the DDSCAT code by Draine & Flatau (1994). The corresponding paper by Draine & Weingartner (1996) clearly demonstrated the ability of the radiation to spin up grains to high rotational velocities. For many researchers in the community, this already appeared to be the solution of the long-standing problem of grain alignment. Indeed, one might think that if the grains rotate fast, they are difficult to randomize and the paramagnetic relaxation should perfectly align such suprathermally rotating grains (Purcell 1979). However, the actual situation is more complicated. The study in Draine & Weingartner (1997) showed that in realistic settings, the grains experience anisotropic radiation flux that makes the grain dynamics far from trivial. Instead of becoming perfectly aligned, grains were shown to exhibit complex phase-space dynamics in which the effects of paramagnetic relaxation torques are negligible compared to the much more powerful torques arising from radiation. The authors did not consider the physics of crossovers (Spitzer & McGlynn 1979), and therefore the modeling was not able to reproduce the actual grain dynamics. These works nevertheless established the alignment of grains by radiation as a promising candidate for explaining how grains can be aligned in interstellar conditions. The actual physics of alignment remained unclear.

A decade later, after a plethora of studies regarding alignment theory, Lazarian & Hoang (2007, henceforth LH07) provided a physically motivated model of the process that successfully explained the properties of grain alignment that were observed in earlier numerical studies. It also corrected mistakes and misinterpretations in these studies. While appreciating the complexity the alignment of realistic irregular grains, the authors identified the radiative torque (RAT) alignment of grains with irregular grains being helical proposed a toy model of a helical grain that allowed analytical description of the alignment. The corresponding analytic model (AMO) of RAT alignment was intended to address the question why radiation tends to align grains with long axes perpendicular to the magnetic field regardless of the direction between the direction of radiation and the magnetic field. The latter contradicted what was suggested in Dolginov & Mytrophanov (1976a). In addition, LH07 introduced a new type of RAT alignment, i.e., the alignment with respect to the radiation direction rather than to the magnetic field. They also provided the criterion for this alignment to take place in the intense radiation flows.

To provide the grain with the property of helicity, LH07 attached a mirror to an oblate grain at 45 degrees and explored the dynamics of this toy system when subjected to radiation flux. 6 The model was surprisingly successful in reproducing not only the general behavior of the irregular grains, but quantitatively reproduced the features of the RAT alignment that were established numerically. In fact, the predictions of the model were successfully compared with the DDSCAT calculations performed for a limited sample of irregular grains available for testing.

Formulation of AMO included identification and analysis of certain key properties of the RAT alignment. Specifically, in the RAT analysis, the RAT is split into three components, out of which the third induces precession of the grain around the radiation field. The ratio of the other two components, qmax, was identified by LH07 to be essential in RAT theory. The quantity qmax was found to depend on the shape of the grain, on the optical wavelength-dependent properties of the grain material, and on the ratio of the radiation wavelength to the effective grain size.

The subsequent studies made use of the AMO, e.g., Lazarian & Hoang (2008, 2019), Hoang & Lazarian (2008, 2009, 2016), and they clarified many essential physical processes of the RAT alignment theory. They confirmed the RAT alignment as the dominant process of alignment for grains in various astrophysical environments. These include the diffuse interstellar medium (ISM), molecular clouds, photodissociation regions, circumstellar regions, and comet comas (see Lazarian 2007; Hoang et al 2015; Kolokolova et al 2016; Tazaki et al 2017).

With the progress of the predictive RAT theory and its successful testing (Andersson et al. 2015), the significance of the RAT mechanism of grain alignment has been generally accepted. However, the regular deviations from the AMO at different wavelengths were observed within the LH07 study using a limited sample of shapes and wavelengths. To deal with this issue, LH07 proposed modifications of the AMO, e.g., for the UV wavelength range. More detailed modifications required studies of more samples, which were not available. It was also shown in LH07 that the degree of alignment depends on qmax, which is a function of several parameters, the grain shape being the main parameter. The earlier studies of qmax were limited to a handful of irregular shapes, and Herranen et al. (2019) first analyzed hundreds of grain shapes. The study opened the way for predicting the actual grain alignment for the ensemble of realistic irregular grains, which is we perform in the present paper.

In Section 2 the grain-alignment problem is formulated. In Section 3 the grain model used in the numerical methods is introduced. In Section 4 the equations of motion relevant in alignment theory are reviewed. Then, DG and RAT alignment mechanisms are compared using RAT results for ensembles of irregular grains. Finally, the fractions of the ensembles for which high-J attractors exist, fhighJ , are determined using a T-matrix method based on a numerically exact volume-integral equation (see Waterman 1965 and Markkanen & Yuffa 2017 for further details). The implications are discussed in Section 5 and results are summarized in Section 6.

2. Formulation of the Problem

The description of grain alignment can be performed in phase space, as was first demonstrated in Draine & Weingartner (1997). It is advantageous to analyze the effects of RATs on grains in the space of angular momentum - alignment angle, where the alignment angle is measured between the angular momentum and either the radiation direction or the external magnetic field direction. In the defined phase space, one can find stationary points. Using this approach, LH07 provided the study for the AMO and defined, as we discuss below, the parameter space for which the expected grain alignment is perfect, i.e., the grain long axes are perpendicular to the magnetic field. This alignment corresponds to grains in stable stationary points (attractors), which correspond to high angular momentum, or high-J attractor points. Note that within the RAT mechanism, this alignment is possible without any assistance from the Davis & Greenstein (1951) paramagnetic relaxation.

If one introduces the measure of alignment of grain angular momentum

Equation (1)

where ξ is the angle between angular momentum and magnetic field direction, then for high-J attractor points, QJ,highJ = 1. For fast-rotating grains, the internal relaxation aligns the angular momentum J with the axis of the maximum moment of inertia X (Purcell 1979). The internal relaxation can arise from the traditional inelastic relaxation (Purcell 1979; Lazarian & Efroimsky 1999), the process called Barnett relaxation in Purcell (1979), or even more exotic, but even more powerful, from the nuclear relaxation introduced in Lazarian & Draine (1999). For typical interstellar conditions, these processes perfectly align J and X, and therefore grains rotate with their long axes perpendicular to the magnetic field B. The corresponding measure of the alignment of grain axes QX,highJ therefore coincides with QJ,highJ , i.e., QX,highJ = 1.

For grains in low-J attractor points, the alignment of J is not perfect because the subthermal angular momentum can easily be randomized by gaseous bombardment. In addition, the thermal fluctuations within the grains induce variations in the direction of X in relation to J (Lazarian 1994; Lazarian & Roberge 1997; Lazarian & Draine 1999). Here, we adopt the value QX,lowJ ≈ 0.25 that was obtained in numerical simulations in Hoang & Lazarian (2008).

As a result, the ensemble of grains that has both low-J and high-J attractor points has the total alignment measure (i.e., Rayleigh reduction factor)

Equation (2)

where fhighJ is the fraction of grains that are aligned in the high-J attractor points. This equation assumes that external and internal alignment are independent of each other, and it is therefore is regarded as a first approximation. The overall measure of the alignment for an ensemble of grains mainly depends on fhighJ .

As an important distinction to other current research, Equation (2) implicitly assumes that all grains for which a high-J attractor exists will be in a high-J state. In steady conditions, collitional excitations will eventually transport grains to a high-J state (Hoang & Lazarian 2008, 2016). However, in recent research (Lazarian & Hoang 2021), an additional high-J criterion has been introduced. The so-called fhighJ,orientation parameter describes the fraction of grains reaching the high-J state quickly, describing fast alignment (Lazarian & Hoang 2007). The fhighJ,orientation parameter is relevant, e.g., in changing environments, where initially randomly oriented particles may not instantly occupy the high-J state, but require time and interactions to eventually reach the high-J state for grains not in the fraction fhighJ,orientation, which reach the high-J state fast. In the current study, fhighJ describes the eventual situation where all those grains are in the high-J state for which this is at all possible.

The existence of high-J attractors depends on the grain interaction with the RATs. These torques in LH07 were decomposed into three components Γ1, Γ2, and Γ3, where the last component of the torque is responsible for grain precession, while the two other components are responsible for the alignment. In fact, it was demonstrated by LH07 that the ratio of the maximum amplitudes of these torques, i.e., ${q}_{\max }={{\rm{\Gamma }}}_{1,\max }/{{\rm{\Gamma }}}_{2,\max }$, is a key parameter that determines the dynamics of grain alignment. This parameter depends on the grain composition, and similar to the grain cross-section, on the ratio of the wavelength to the effective grain size λ/a.

The LH07 study introduced the requirements for the existence of the high-J attractors, ${q}_{\max }$, and the angle ψ between the direction of radiation and the magnetic field. For the parameter space in which there are no high-J attractor points, the RATs decrease the grain rotational velocities, bringing grains in low-J attractor points.

While the RAT alignment does not require any effect of paramagnetic relaxation, it was demonstrated in Lazarian & Hoang (2007) that an enhanced magnetic dissipation can stabilize high-J stationary points, transferring them to high-J attractor points. The effect was quantified in Hoang & Lazarian (2016), with the results reproduced in Figure 1 using an idealized 7 AMO of a spinning grain with torque components (LH07)

Equation (3)

Above, the angle Θ in scattering frame can be written in terms of the alignment frame angles ξ, ψ, and ϕ (see Figure 2) as (Draine & Weingartner 1997)

Equation (4)

Figure 1.

Figure 1. Contour map of the critical magnetic inclusion ratio δm required for AMO to produce high-J attractor points with respect to the q-factor qmax and angle ψ between radiation and the magnetic field directions.

Standard image High-resolution image
Figure 2.

Figure 2. Alignment coordinate system.

Standard image High-resolution image

Additionally, a factor

Equation (5)

is introduced, where τran is the time of randomization by different processes that include gas bombardment, emission of photons, and interaction with ions (see Draine & Lazarian 1998, for a complete list of randomizing processes), and τm is the time of the magnetic dissipation of the rotational energy of the grain. The parameter δm is a key factor for the Davis-Greenstein mechanism, but for the RAT alignment, it plays an auxiliary role in increasing fhighJ .

While the angle ψ is determined by the relative position of the radiation source to the direction of the grain precession about the magnetic field, ${q}_{\max }$ is an intrinsic property of the grain that is determined by its shape and composition. Our work intends to determine the percentage of realistic irregular grains that can have high-J attractor points for different settings.

3. Sample of Grains to Study

We analyzed the RATs on irregular grains using ensembles of Gaussian random ellipsoids (GRE; Muinonen & Pieniluoma 2011). A single GRE shape can be identified by two statistical parameters: the correlation length between displacements, and the standard deviation of displacements from the undeformed ellipsoid surface. Regeneration of a single explicit shape is possible by fixing the axial ratios of an ellipsoid and resetting the seed of a random number generation. The method provides the means for producing large numbers of random shapes, suitable for assessing the predictions of AMO, although GREs are unlikely to be a perfect representative of interstellar grain shapes.

We consider ensembles of oblate and prolate grains of three different sizes, with equivolume spherical radii aeff of 0.1 μm (N = 1000), 0.2 μm (N = 1000), and 0.75 μm (N = 400). The grains are composed of astronomical silicate (Draine & Lee 1984), which has a complex refractive index close to n = 1.68 + i0.03 in the visual wavelength range.

The grain shapes are deformed from a smooth oblate/prolate spheroid to Gaussian random oblate/prolate spheroids (hereafter referred to as Gaussian oblate/prolate grains). The deformations for the basic shape, which have aspect ratios a: b: c = 1: 1: 0.5 for the oblate shape and 1:0.5:0.5 for the prolate shape, all follow lognormal statistics for the radius deformation with a standard deviation 0.125, and a correlation length 0.35 between points on the spheroid.

The Gaussian random deformation process provides means of studying statistical samples of random shapes with control over their general inertial properties using a minimum amount of parameters. The chosen aspect ratios reflect the assumption that high rotational speeds would likely disrupt (Hoang et al. 2019; Hirashita & Hoang 2020) more extreme shapes 8 and that sphere-like inertial and shape properties are highly unlikely to differ from results obtained by a Mie scattering analysis of spheroidal grains.

For the sake of consistency, the statistical behaviors of the deformed grain radii and moments of inertia are studied. The principal moments of inertia for an oblate and a prolate spheroid are

Equation (6)

where diag denotes elements on a matrix diagonal. We compare the statistical average and standard deviation of diagonalized inertia matrix components to determine whether the inertial properties of the base shape are conserved. We also compare the deformed volumetric radii Rvol (also the radius of the sphere of equivalent volume, or aeff) with those of the base shapes. A distinction between the two is made here because in the scattering calculations all the grains are scaled so that a desired aeff is achieved, whereas here we use Rvol to measure with a single number whether the shape deformation affects the overall shape. For a general ellipsoid, the volumetric radius is simply Rvol = (abc)1/3. The results are collected in Table 1. We see that the randomization process indeed produces the expected values, which are close to the value of the base shapes.

Table 1. Ensemble Properties Compared to the Undeformed Base Shapes

  Rvol E[Rvol] σ diag (Ip ) E[diag (Ip )] σ
Oblate0.79370.80260.0482(5,5,8)(4.185, 5.899, 8)(0.355, 1.183, 1.101)
Prolate0.63000.64540.0582(2,5,5)(1.978, 4.578, 5)(0.306, 1.009, 0.954)

Note. The dimensionless units of length are chosen so that the semimajor axes have values 1 and the semiminor axes have values 0.5. The components of Ip and the corresponding standard deviations have been scaled so that the last components, or maximum moments of inertia, are equal. The correlation length is fixed at 0.35 for all shapes.

Download table as:  ASCIITypeset image

4. Alignment Fraction Analysis for Gaussian Random Ensembles

With the GRE ensemble as a starting point, normalized radiative torques are solved using the T-matrix method. As the T-matrix describes grain scattering properties at a single wavelength, the following results are calculated for a 10-element sample of a wavelength range λ ∈ [300, 1920] nm, equally divided between the endpoints.

For maximum applicability of the results in astronomical context, the interstellar radiation field (ISRF) of the local solar neighborhood (Mathis et al. 1983) is hereafter assumed as the distribution of incident energy density, unless explicitly wavelength-dependent quantities are considered. When RATs under ISRF illumination are considered, they are computed using the definition

Equation (7)

4.1. Equations of Motion for Alignment

We work in the alignment coordinates (ξ, ψ, ϕ), where ξ is the alignment angle between angular momentum and the magnetic field vectors J and B , ψ is the angle between B and the anisotropic radiation direction k , and ϕ is the Larmor precession angle of J around B . The coordinate system is illustrated in Figure 2. In the following analysis, averaging a quantity q over ϕ is denoted by $\bar{q}$.

Using the equations of motion of RATs where perfect internal alignment is assumed, we have, similarly as in LH07, for the alignment angle ξ and angular momentum $J^{\prime} =J/{I}_{1}{\omega }_{T}$ (Hoang & Lazarian 2016)

Equation (8)

Equation (9)

Above, $t^{\prime} =t/{\tau }_{\mathrm{gas}}$, I1 is the maximum grain moment of inertia, ${\omega }_{T}={(2{k}_{B}{T}_{\mathrm{gas}}/{I}_{1})}^{1/2}$, $M=\gamma \bar{\lambda }{u}_{\mathrm{rad}}{a}_{\mathrm{eff}}^{2}/2$. Finally, $\bar{F}$ and $\bar{H}$ are the ϕ-averaged aligning and spin-up RAT components.

The stationary points (ξs , Js ) can be found by setting the equations of motion to zero, which leaves us with

Equation (10)

The zeros of Equation (10) now give, as shown in LH07, the stationary points in the (ξ, J) phase space. Specifically, we find universal stationary points $\sin {\xi }_{s}=0$, which have, after ϕ-averaging, $\bar{F}=0$.

Generally, for a stationary point to be an attractor, we have the requirements, as given by Draine & Weingartner (1997; DW97),

Equation (11)

where

Equation (12)

and ${\rm{\nabla }}\equiv \tfrac{{\rm{d}}}{{\rm{d}}\xi }$. When we consider the universal stationary points, the requirement for high-J attractors simplifies to

Equation (13)

When the equations of motion are analyzed for grain ensembles as a function of ψ, we can find fhighJ . In the following calculations, we assume that the maximum axes of inertia of grains are aligned with the magnetic field, i.e., the universal stationary points alone are considered. Physically, this corresponds to a situation in which the alignment process has reached a steady state, and all alignable grains are aligned. Thus, the fraction fhighJ is the fraction of grains for which requirement (13) for any of the two universal stationary points is satisfied.

4.2. Distributions of qmax of GRE Grains

In order to derive predictions for alignment degrees with AMO, a distribution of qmax corresponding to some dust grain population is required. Now, we focus on qmax distributions obtained from the GRE ensemble. In Figure 3 we present distributions of qmax for the three differently sized oblate and prolate GREs, along with lognormal fits for the distributions.

Figure 3.

Figure 3. Distributions of qmax values along with lognormal fits for each subensemble of separate size and base shape.

Standard image High-resolution image

We chose a lognormal distribution to provide the fits for the adequate quality of the fit and because the generation of GRE shapes involve lognormal statistics. It should be noted that no claims on the connection between the shape of the qmax distribution and the underlying GRE statistics are made, and such an analysis is well beyond our current scope.

The lognormal distribution, which can be generated using logmean μ and logdeviation σ as parameters, provides adequate fits for all the separate-size-shape grain populations. Furthermore, for the total ensemble of 2400 grains, the lognormal distribution provides a good fit with σ = 0.37, μ = 1.17 (see Figure 10, right panel).

The mean value of the subensemble qmax distribution increases with the grain size in the ISRF. For extrapolations, it is important to identify the functional form of the average qmax with respect to the ratio λ/aeff of the wavelength and grain size. In Figure 4 we present the mean values for the oblate and prolate subensembles.

Figure 4.

Figure 4. Average ${\bar{q}}_{\max }$ as a function of λ/aeff. Vertical bars represent one standard deviation.

Standard image High-resolution image

We identify a significant peak in ${\bar{q}}_{\max }$ (bar indicating an ensemble mean) at λ/aeff = 2 for both the prolate and oblate subensembles. In terms of grain sizes, there thus will be some grain size for any given incident energy density distribution for which a maximum ${\bar{q}}_{\max }$ is obtained. Now, the ISRF according to Mathis et al. (1983) peaks at 700 nm, and the UV peak at 120 nm is left outside the wavelength range considered.

When we extrapolate a similar functional form of ${\bar{q}}_{\max }$ for any grain size, the peak location will correspond to a grain size of 0.35 μm. For the ISRF, the distribution of qmax for grains with aeff = 0.35 μm would therefore likely tend most toward high values.

Additionally, we note that ${\bar{q}}_{\max }\approx 1$ is a local minimum between λ/aeff = 4 and 5 with a low standard deviation, especially in the oblate case. At this size range, we therefore expect the ensemble behavior to be one of the most representative behaviors of the difference between irregular grains and AMO, which predicts no high-J attractors without magnetic inclusions at ${q}_{\max }=1$.

The distributions of qmax have mean values close to ${q}_{\max }=3$. Considering AMO, using Figure 1, we predict that if most grains have qmax values close to 2 or higher, the fraction fhighJ will quickly decrease to zero at around ψ = 30°. Next, we test how the GRE ensemble results compare to the AMO prediction.

4.3. Ensemble high-J Fractions fhighJ

We compute the high-J alignment fractions for the GRE subensembles using the procedure described in Section 4.1. For simplicity, magnetic inclusions are not considered, i.e., δm = 0.

First, in the left panel of Figure 5, the fhighJ of the subensembles are compared with the prediction of AMO with a lognormal (σ = 0.37, μ = 1.17) qmax distribution. As δm = 0, the fhighJ of AMO falls to zero just after ψ has reached 30°, as is predictable from Figure 1 for a distribution of qmax that does not contain grains with ${q}_{\max }\lt 0.5$.

Figure 5.

Figure 5. Left panel: Fraction fhighJ of grains with high angular momentum attractors of the six different GRE subensembles and AMO as a function of the angle ψ between the magnetic field and the radiation field directions. Right panel: Similar to the left panel, but for a reorganization of the ensembles according to grain size and the qmax value compared to the average qmax of each different grain sizes.

Standard image High-resolution image

The GRE ensembles exhibit high-J attractors more universally than predicted by AMO. Two important observations can immediately be identified from the data: First, fhighJ both increases and becomes more uniform with respect to ψ along with increasing grain size. Second, for all subensembles, the minimum value of fhighJ is centered around ψ = 40°, near the angular range for which AMO predicts no high-J attractors without paramagnetic inclusions, and the smaller the grain, the wider the minimum. Most grain types also exhibit a local maximum around ψ = 70°–80°.

The difference between fhighJ of the oblate and prolate grains is slightly in favor of oblate grains. The difference between base shape types is mostly smaller than 10 %, with the exception of the 0.2 μm cases in the approximate range ψ ∈ (60°, 70°).

Second, in the right panel of Figure 5, a comparison of rearranged subensembles is performed. The rearrangement is done so that grains of a certain size with qmax values lower or higher than the mean qmax value of all grains at this size are separated into different subensembles.

According to Figure 1, grains with large qmax are expected to have higher fhighJ at small ψ and lower fhighJ large ψ when compared to the counterpart with low qmax values. However, we see that this is not the case for irregular grains. Instead, grains with larger qmax have almost universally at least as large fhighJ as the counterpart with smaller qmax, the most notable exception being the 0.75 μm case around ψ = 40°.

Third, in order to identify the contribution of RATs by different wavelengths on fhighJ , Figures 6 and 7 show fhighJ as a function of λ/aeff and ψ.

Figure 6.

Figure 6. Fraction fhighJ as a function of λ/aeff and ψ for the three oblate subensembles. All panels share the same color scaling.

Standard image High-resolution image
Figure 7.

Figure 7. Same as Figure 6 for the three prolate subensembles.

Standard image High-resolution image

Comparing the columns of Figures 6 and 7 to the functional form of ${\bar{q}}_{\max }$ confirms that a high value of qmax is required to produce high-J attractors at small ψ. Furthermore, as ${\bar{q}}_{\max }$ of both the oblate and prolate cases are highly similar, so are Figures 6 and 7.

Next, Figures 6 and 7 can be compared against Figure 1 with the help of ${\bar{q}}_{\max }$ data from Figure 4. At λ/aeff ≈ 5, where ${\bar{q}}_{\max }$ is at a minimum, the contour structure of Figure 1 predicts that the cutoff angle ψ for which high-J begins to occur should also be at a minimum. This slight drift is visible in both Figures 6 and 7. Additionally, in the middle of the cross-shaped contour structure, no high-J attractors should exist without magnetic inclusions. For AMO, the high-J poor region occurs at ${q}_{\max }\approx 1.2$, while for irregular grains, the region of the least high-J attractors occurs at λ/aeff ≈ 3, at ${q}_{\max }\approx 2$, or at 3 (for small prolates).

As a last observation, the 0.75 μm cases clearly show the discrepancy between fhighJ of AMO and the GRE ensemble for ψ < 40°. As shown in Figure 4, only a small fraction of the grains have ${q}_{\max }\lt 1$, but more high-J attractor points exist at values of ψ higher than 40°.

It is important to note that because qmax is not a free parameter for irregular grains and because the irregular grain radiative torques can have a highly variable functional form for two grains with similar qmax, a figure such as Figure 1 is difficult if not impossible to produce for the GRE ensemble. Thus, a comparison of the radiative torques of AMO and of the GRE ensemble is crucial in order to explain the discrepancy between fhighJ of AMO and the GRE ensemble of Figure 5.

4.4. Differences between the GRE Ensembles and AMO

Potentially, the main explanation for the differences between obtained GRE ensemble results and the predictions derived from AMO is the different functional forms of the RATs themselves. In this section, we compare the RAT components in both the scattering frame and in the alignment frame. In both cases, the mean square differences $\left\langle {{\rm{\Delta }}}^{2}\right\rangle q$ of the GRE and AMO quantities q,

Equation (14)

are determined.

First, we compare in Figure 8 the shapes of the scattering frame RAT components Γ1 and Γ2. In the procedure, the AMO RAT is produced according to the qmax of each grain in the GRE ensemble and by normalizing the torques with the maximum values of Γ1. The mean square difference is then calculated over the angle Θ, and the result is presented as a function of λ/aeff.

Figure 8.

Figure 8. Mean square differences over the angle Θ of the scattering frame RAT components Γ1 and Γ2 between the AMO and the GRE subensembles as a function of λ/aeff.

Standard image High-resolution image

According to Figure 8, the AMO is generally more likely to produce similar RATs to GRE grains the smaller the grains are. The result is consistent with the results of Figures 6 and 7, which show that smaller grains exhibit fhighJ values closer to those predicted by the AMO.

Additionally, the difference $\left\langle {{\rm{\Delta }}}^{2}\right\rangle {{\rm{\Gamma }}}_{2}$ reaches a peak value around λ/aeff = 2. The difference for Γ2 remains large for larger grains, and for Γ1, it increases scrictly as the grains grow larger. For large grains, the predictions of the AMO can therefore deemed to be relatively inaccurate.

Next, alignment frame RATs are compared in order to potentially identify critical differences between the AMO prediction of Figure 1 and fhighJ of the GRE subensemble. The differences are determined over the whole GRE ensemble as a function of the angle ψ. Additionally, the sign difference percentages of ∇F between the AMO and the GRE subensembles are determined. The results are collected in Figure 9.

Figure 9.

Figure 9. Mean square differences in the alignment frame over λ/aeff of ∇F (top panel) and H (bottom panel) between the AMO and the GRE subensemble as a function of the external alignment angle ψ between the magnetic field and the radiation directions. Middle panel: The percentage of GRE subensembles with a different sign of ∇F at $\sin \xi =0$ as a function of ψ.

Standard image High-resolution image

Irregular grain RATs, particularly ∇F and its sign, differ from those of AMO at high values of ψ. The spin-up component H appears to approach that of the AMO as ψ approaches 90°.

Differences of ∇F are locally very small at values around ψ = 40°, where fhighJ reaches the minimum value with both the AMO and the GRE subensembles (Figure 5, left panel).

The sign difference of the AMO and the GRE subensembles reaches up to 50% at ψ = 90°. However, fhighJ tends to decrease at ψ > 80° for all grains to as low as 0.2, implying that the attractors in this region are low high-J ones.

To summarize the section, the RAT differences show an expected size dependence that is evident from the fhighJ results. Particularly the differences in the scattering frame component Γ2, which is expected to be important near ψ = 90° (where the radiation direction is perpendicular to J ), and the differences of the ξ-derivative of aligning the RAT component F are likely to be related to high fhighJ values that are not predicted by the AMO at large ψ.

4.5. Effect of Magnetic Inclusions

According to Figure 1, high-J attractors are universal with moderate magnetic inclusions, expect in the limit of large qmax and ψ. We test the existence of high-J attractors when magnetic inclusions are added for the total GRE ensemble in Figure 10.

When δm is increased to 10, this increases fhighJ to 0.9 when ψ < 70°, i.e., high-J attractors become universal at ψ < 70°. Above ψ = 70°, fhighJ decreases to 0.5, as seen in the left panel of Figure 10.

Figure 10.

Figure 10. Left panel: Values of fhighJ as a function of ψ for different values of δm describing the different relative importances of magnetic inclusions. We plot an AMO with δm = 0 for reference. Right panel: Distribution of qmax for the total GRE ensemble.

Standard image High-resolution image

Figure 1 predicts that at ψ = 90°, for grains with ${q}_{\max }\gt 2.5$, high-J attractors do not exist without extreme magnetic inclusions, if at all. According to the right panel of Figure 10, the mean value of ${q}_{\max }=2.762$. Thus, approximately half of the grains have ${q}_{\max }\gt 2.5$, which makes the result in the left panel of Figure 10 consistent with Figure 1 and also leads to the conclusion that in the limit of large qmax and ψ, high-J attractors do not exist.

4.6. Total Degree of Alignment

The total degree of alignment is approximately given by Equation (2), and it depends on the degree of internally aligned grains with either low-J or high-J attractors. We recall that for grains with a low-J attractor only, the degree of internal alignment was chosen at QX,lowJ = 0.25. This results in the total degrees of alignment presented in Figure 11.

Figure 11.

Figure 11. Total degree of alignment as a function of ψ for the GRE subensemble of different sizes, and for an AMO with a lognormal qmax distribution.

Standard image High-resolution image

The resulting degrees of alignment, where prolate and oblate subensembles are combined, resemble the fhighJ results from Figure 5, as expected. The choice of QX,lowJ guarantees that AMO at ψ > 30° gives the lowest possible degree R = 0.25. However, especially the large 0.75 μm grains clearly have a higher degree of alignment.

In the approximative scheme, where the total degrees of alignment are derived from Equation (2), fhighJ provides a straightforward means to solve for R with any assumption on QX,lowJ/highJ values.

5. Discussion

5.1. Main Points of this Study

LH07 identified the grain helicity as the driver of the RAT alignment and presented a toy model of a helical grain that allowed a simple analytical description, i.e., the model of the AMO. A significant progress in understanding the properties of grain alignment was achieved by the AMO. Nevertheless, one would not expect the toy model to perfectly represent the RAT alignment of an ensemble of irregular grains. In fact, deviations of the RAT functional form from AMO predictions were reported in LH07. 9 However, with just a few shapes studied there, it was not possible to evaluate the consequences of these deviations on the degree of alignment achievable by an ensemble of irregular grains. The present paper provides the statistical data needed by using an extensive number of irregular grains subjected to the interstellar radiation field.

The polarization predicted by the AMO was insufficient to explain the maximum values of polarization observed in the ISM, at least in the dust model where only silicate grains are aligned. 10 This made the AMO model with enhanced magnetic dissipation, introduced in Lazarian & Hoang (2008), relevant. The detailed predictions of the AMO were presented in Hoang & Lazarian (2016, HL16).

An important point of the present study is that our results show a higher alignment efficiency than for the AMO, opening the possibility that the observed grain alignment can be accounted for by the RAT alignment of ordinary paramagnetic grains. However, magnetic inclusions cannot be immediately disregarded based on our results, where R ∼ 0.4–0.5 for grain sizes <0.3 μm. Recently, an analysis of spectral polarization by Draine & Hensley (2021) required R ∼ 0.7 to explain the 10 μm polarization feature, assuming a reasonable axis ratio of spheroidal grains (between 1/3 and 3, favoring more modest ratios). The axis ratios of the undeformed spheroids in this study were 0.5 and 2. However, the higher alignment degrees of irregular grains, compared to AMO, allow the possibility that in some cases, the RAT alignment degrees of irregular grains may be sufficient to explain some observations even if grains do not have magnetic inclusions.

HL16 presented a contour map for the critical magnetic relaxation parameter, δm,cri for AMO (adapted in Figure 1), required to produce high-J attractor points. There, it was found that for δm > 10, high-J attractors are universal, and for lower values, there are regions of interest both in terms of qmax and alignment angle ψ.

As a rule of thumb, the qmax values near unity are most problematic for the existence of high-J attractors. HL16 found that when ψ = 45° is approached from smaller angles, δm,cri is smallest for grains with qmax smaller than unity. The same is true for qmax larger than unity when it is approached from the large angle side. This implies that fhighJ values should drop for numerical ensembles, given that the distribution of the qmax factor contains values near unity, which was systematically observed in Section 3. The exact value for the δm,cri and the critical value for the qmax likely differ from those obtained from the AMO, but the trends suggested by the AMO nevertheless appear.

In our wavelength-dependent study, we find that as grains become smaller, qmax generally tends to decrease gradually, although the special significant drop around λ/aeff = 5 introduces another general rule to the shape of ${q}_{\max }$ as a function of λ/aeff. Thus, the decrease in fhighJ as a function of ψ (Figures 6 and 7, 0.1 μm and 0.2 μm cases) should slightly drift toward lower values of ψ in this size range. The effect is produced by our study, but the drift is very subtle and agrees well with the results predicted using the AMO.

In the leftmost columns, the 0.1 μm and 0.2 μm cases in Figures 6 and 7, the values of fhighJ become more uniform as ψ > 50°. The size region is probed in more detail for 0.75 μm grains (N = 400). If qmax and fhighJ are connected, the spread of qmax values as grains grow results in fhighJ values that can be explained by the AMO prediction from HL16. HL16 states that for qmax values greater than unity at 0° < ψ < 45°, grains with large qmax should be well aligned, and again, this is similar for ψ > 45° and small qmax. Here, the idealized AMO predicts similar results, but for grains with large qmax, the upper limit of the range of well-aligned particles is reduced to about ψ = 35°. As the spread of the qmax distribution is not extreme for large grains, the more uniform fhighJ must also be attributed to the RAT shape differences between the AMO and irregular grains.

Using numerical evidence of ensembles of irregular grains, it was found that AMO can successfully reproduce the basic features of the RAT alignment observed with the large sample of GRE grains. In particular, this explains why the alignment takes place with long axes perpendicular to magnetic field, it explains that the alignment can occur at low-J and high-J attractor points, and it explains the alignment properties that are achieved at these points. It also explains the decreasing maximum rotational rate with increasing ψ, which becomes rapid as ψ exceeds 60°. For irregular grains, this decrease is observed around ψ ∼ 45°, beyond which fhighJ fluctuates. The fluctuations in the grain-ensemble-averaged fhighJ can likely be attributed to the irregular shape itself, due to which ϕ-averaged RATs at certain angles provide attractor points.

The importance of qmax was further solidified by the evidence because it correlates with the likelihood of existence of high-J attractor points. The numerical results also imply that while the exact point of critical values for the parameters may differ for different types of grains due to the complicated RAT shapes of irregular grains, the general behavior is also correctly produced by the AMO.

At the same time, our results show the limitations of the AMO in terms of a qualitative description of the properties of some grains in the ensemble of shapes we studied. We will address elsewhere to what extent this is limited to the shapes of higher elongation and whether the correspondence can be obtained with a modified AMO, e.g., by an AMO with a mirror turned at a different angle, as suggested as a modification in LH07. The advantage of the AMO is its extreme simplicity, and having a set of different modifications might not be convenient. The development that we find very positive is that the RATs acting on the studied ensemble of grains can provide a higher alignment degree than the classical model of the AMO for large angles between the radiation direction and the magnetic field (i.e., ψ > 45°, see Figure 11). This gives us hope that the observed interstellar alignment can be explained without evoking an enhanced magnetic susceptibility of grains.

5.2. Role of the RATs and Limitations of the Present Study

As we mentioned above, the most important question addressed in this paper is whether the enhanced magnetic dissipation is necessary to explain high degrees of alignment. The LH07 theory was formulated for no additional relaxation, while it was shown in Lazarian & Hoang (2008) that higher degrees of alignment can be achieved if the high-J stationary points are stabilized in the presence of a strong magnetic response of the grains. The arguments in favor of the magnetic grains are presented in Hoang & Lazarian (2016), while Lazarian & Hoang (2019) discussed the observational procedures for quantifying the grain magnetic properties.

Our present work opens up the discussion about the composition of interstellar grains. An important question is whether magnetic inclusions (Lazarian & Hoang 2008; Hoang & Lazarian 2016) are necessary to explain the observed interstellar alignment. The increased grain-alignment efficiency reported in this study facilitates explaining interstellar grain alignment. The task becomes even more simple if a significant part of interstellar carbon is locked in the composite grains consisting of carbonaceous and silicate fragments. We note, however, that our ability to explain the interstellar polarization with paramagnetic grains does not actually mean that the grains do not contain magnetic inclusions.

For instance, most of the interstellar iron is locked in grain material, but we do not know of which chemical substance iron is a part and what the magnetic properties of the substance are. Thus, it is important to test the composition of grains by measuring the polarization and its changes in special settings, for instance, near stars of transient radiation sources, e.g., novae. Furthermore, existing studies have constrained the possibility of magnetic inclusions by comparing predictions (Draine & Hensley 2013) with observations (Planck Collaboration et al. 2020). The corresponding tests of the magnetic properties of grains based on the grain-alignment theory are discussed in Lazarian & Hoang (2019), while in Lazarian & Hoang (2021), the tests related to the alignment-dependent grain disruption are proposed.

In the present paper we considered the B-alignment, i.e., the alignment with respect to the magnetic field. As was demonstrated in LH07, the alignment can also occur with respect to the direction of radiation, provided that the precession induced by the Γ3 component of the radiative torque is faster than the grain precession about the ambient magnetic field. The process of k-alignment can be analyzed with our present approach by assuming ψ = 0.

The RAT alignment process is complex, with only a small fraction of grains initially reaching the high-J attractor points. In the presence of time-dependent radiation sources, LH07 demonstrated that transient alignment takes place. During the process, most grains are aligned in the low-J attractor point on the timescale of the order of grain precession induced by the Γ3 RAT component. However, on the time of several τran, the grains reach stationary alignment at high-J attractor points if such points are present for the given combination of qmax and ψ.

For our calculation in the paper we assumed that the grains are subject to the sufficiently strong radiation field that causes grains at high-J attractor points to rotate with velocities significantly faster than the thermal velocity. These conditions are well satisfied in the diffuse ISM for ∼0.1 μm grains. For grains shielded from the radiation in molecular clouds or accretion disks, as well as for grains with sizes appreciably different from the typical wavelength of the impinging radiation, the rotational rates of grains at high-J points can be reduced to become comparable to the thermal rotation speed. Such grains can be randomized by gaseous collisions and other randomizing factors. The RAT alignment of these grains in these conditions is decreased for ψ larger than 80 degrees. According to Figure 8 in Lazarian & Hoang (2019), the RAT strength decreases for this range.

The current study is for the interstellar radiation field. The ISM is the area in which the grain alignment has been studied most frequently (see Andersson et al. 2015). However, the RAT alignment of dust is a ubiquitous process that can take place in different astrophysical environments and in planet atmospheres (see Lazarian 2007). Therefore detailed studies of the peculiarities of grain alignment for different spectra of radiation-illuminating ensembles of irregular grains of different compositions are required. This work can therefore be viewed as the first step in this direction.

5.3. Application to Carbonaceous Grains

The properties of RATs we explored in this paper are applicable for grains made of different materials. The change in optical constants is not expected to radically change the grain alignment. On the basis of the obtained results, we can therefore consider the alignment of not only silicate, but also of carbonaceous grains. The RAT alignment of carbonaceous grains was considered in Lazarian (2020), and it was shown that carbonaceous grains can exhibit a different mode of RAT alignment, i.e., the alignment with long grain axes parallel to the magnetic field. The anomalous mode arises due to charged carbonaceous grain precession along the electric field. The latter arises due to the grain motion with respect to magnetic field.

The difference between the alignment of carbonaceous and silicate grains considered in this study arises from the difference of the magnetic moment of the grains. The latter is significantly smaller for carbonaceous grains. Our qualitative conclusions on the modification of grain alignment compared to the AMO prediction remain true, but in terms of observed polarization, the observed degrees can be lower due to grain rotation about the magnetic field.

If carbonaceous grains reach the state of high-J rotation, their magnetic moment increases, and it was shown in Lazarian (2020) that such grains can align with long grain axes perpendicular to the magnetic field, i.e., they align in a similar way to silicate grains. Our present study shows that the percentage of grains in the high-J rotation state for irregular grains is higher than in the AMO prediction. Therefore we may expect more carbonaceous grains to become aligned in the regular fashion, i.e., with long grain axes perpendicular to the magnetic field.

The alignment of grains with long axes parallel to the magnetic field can, as explained in Lazarian (2020), result in a more complicated pattern of polarization that is more difficult to interpret in terms of the underlying magnetic field. At the same time, if actual interstellar grains are composite, i.e., contain both carbonaceous and silicate fragments, their alignment is similar to that of silicate grains.

5.4. Are Highly Irregular Grain Shapes Present in the ISM?

The ensemble of irregular grain shapes used for our study and comparison to AMO is generated using a Gaussian random algorithm, which is not subject to any astrophysical constraints. Therefore a considerable fraction of generated grains has extreme shapes characterized by a high ratio of RAT components, qmax < 0.5 or ${q}^{\max }\gt 5$, which induces a higher degree of alignment for ψ > 45° than predicted by the AMO. However, whether such highly irregular shapes can exist in the ISM remains unclear. Highly irregular shapes are likely to have a low tensile strength, so that they are more easily disrupted by by RATs (Hoang et al. 2019; Hoang 2019, 2020). More observational studies on grain shapes are required to test the RAT alignment and disruption theory as well as the relation of the alignment and the disruption (see Lazarian & Hoang 2021).

6. Summary

In the paper we have explored the grain alignment for a collection of irregular grains of different shapes. We compared the properties of the RAT alignment with the standard AMO. While the deviations in the RAT functional shape from the AMO prediction are not significant, we report a general improvement of the alignment compared to the AMO expectations. On this basis, we may argue that the explanation of the observed alignment can potentially be achieved with ordinary paramagnetic grains without magnetic inclusions. Our results can be briefly summarized as follows:

  • 1.  
    A better alignment of irregular grains by RATs compared to the predictions by the standard AMO is found, especially when alignment at large ψ is considered. In particular, ∼40% of the irregular grains at ψ = 80° deviate from the AMO predictions and exhibit high-J attractors. In this situation, it requires more theoretical and observational research to conclude whether magnetic inclusions are necessary to explain the properties of interstellar polarization.

J.H. thanks the UW Department of Astronomy for its hospitality during his visit. A.L. acknowledges the support of NSF grants AST 1715754, and 1816234. The Flatiron Institute is supported by the Simons Foundation. T.H. acknowledges the support from the National Research Foundation of Korea (NRF) grants funded by the Korean government (MSIT) through the Mid-career Research Program (2019R1A2C1087045). We thank the referee for numerous comments that helped us to improve the presentation of our results.

Note added in proof

Our more recent study shows that, as we mentioned in the paper, the modification of AMO via changing the angle that the mirror makes with the circumference of the grain indeed improves the correspondence of the AMO predictions with the results obtained for the ensemble of grains studied numerically in this paper. In other words, while the original AMO with a mirror attached at $\alpha =45$ degrees provides a good correspondence for many grain shapes, the variations of $\alpha $ allows to us fit the properties of very different grain shapes interacting with electromagnetic radiation at very different wavelength in agreement with the original study in Lazarian & Hoang (2007, LH07).

The combination of the standard AMO (α = 45°) and a modified AMO with the α ≲ 30° would decrease the fraction of high-J attractors at ψ < 45° and increase the fraction of high-J at ψ > 45°. From Figure B16 in LH07 one can see that qmax decreases with decreasing α so we can combine AMO with different angles such that, e.g, ${f}_{45}{\rm{AMO}}(\alpha =45)\,+\,(1-{f}_{45}){\rm{AMO}}(\alpha =20)$ with f45 is the fractions of the grains with the mirror at 45°. Our estimates show that with f45 ∼ 0.7, we can bring AMO to be in agreement with GRE as shown in Figure 5. Naturally, this is just a fit, but it shows the power of AMO in terms of explaining properties of RATs for ensembles of irregular grains.

The important new finding in this paper is that those adjustments increase the parameter space for which one expects perfect RAT alignment. This indicates that the necessity of superparamagnetic enhancement of the RAT alignment (Lazarian & Hoang 2008; Hoang & Lazarian 2016) may not be necessary in order to explain observations. The latter point requires further studies.

Footnotes

  • 6  

    This simple model opened a way to predict a new type of alignment, i.e., to predict that the irregular grains can also be aligned with their long axes perpendicular to the magnetic field by a gaseous flow, and this alignment is similar to the RAT alignment (LH07).

  • 7  

    Meaning the simplest form of equations reproducing torque components similar to those considered in LH07 for a mirror at a 45° angle measured between the mirror normal and the AMO grain intermediate axis; see LH07 for the exact definition.

  • 8  

    The relation between the grain disruption and the grain alignment is explored in Lazarian & Hoang (2021). It is shown there that in some situations, e.g., in accretion disks, grains at high-J state of rotation can survive in the presence of intensive radiation. However, we consider a more general case of interstellar disruption.

  • 9  

    In LH07, modifications of the AMO by changing the angle of the mirror to the grain axis of the maximum moment of inertia were considered as a way to reproduce torques acting for grains at different wavelengths. In the absence of sufficient numerical input, the basic AMO with the mirror at an angle of 45 degrees was used in the subsequent studies (see Hoang & Lazarian 2008, 2009, 2016).

  • 10  

    If the aligned grains are composite, i.e., contain both silicate and carbonaceous fragments, the constraints on the alignment efficiency are reduced. The abundunce of these composite grains is a controversial issue, however.

Please wait… references are loading.
10.3847/1538-4357/abf096