Brought to you by:

The following article is Open access

Inhomogeneity within Local Interstellar Clouds*

, , , and

Published 2022 August 22 © 2022. The Author(s). Published by the American Astronomical Society.
, , Citation Jeffrey L. Linsky et al 2022 AJ 164 106 DOI 10.3847/1538-3881/ac816b

Download Article PDF
DownloadArticle ePub

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

1538-3881/164/3/106

Abstract

Analysis of interstellar absorption lines observed in high-resolution Hubble Space Telescope spectra of nearby stars provides temperatures, turbulent velocities, and kinetic properties of warm interstellar clouds. A new analysis of 97 interstellar-velocity components reveals a wide range of temperatures and turbulent velocities within the Local Interstellar Cloud (LIC) and the nearby Cluster of Interstellar Clouds (CLIC). These variations appear to be random with Gaussian distributions. We find no trends of these properties with stellar distance or angles from the Galactic Center, magnetic field, the main source of extreme-UV radiation (the star epsilon CMa), the center of the LIC, or the direction of inflowing interstellar matter into the heliosphere. The spatial scale for temperature variations in the LIC is likely smaller than 5100 au, a distance that the Sun will traverse in 1000 yr. Essentially all velocity components align with known warm clouds. We find that within 4 pc of the Sun, space is completely filled with partially ionized clouds, but at larger distances space is only partially filled with partially ionized clouds. We find that the neutral hydrogen number density in the LIC and likely other warm clouds in the CLIC is about 0.10 cm−3 rather than the 0.20 cm−3 density that may be representative of only the immediate environment of the LIC. The ≤3000–12,000 K temperature range for the gas is wider than the predictions of thermal equilibrium theoretical models of the warm neutral medium and warm ionized medium, and the high degree of inhomogeneity within clouds argues against simple theoretical models.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The heliosphere does not live in static isolation, but is instead encapsulated by an environment with inhomogeneous spatial properties that manifest as time-variable external properties as the heliosphere traverses the interstellar medium (ISM). Total pressure balance between the outer heliosphere, hereafter called the Very Local Interstellar Medium (VLISM), and the surrounding ISM can greatly alter the size and shape of the heliosphere. For example, Zank & Frisch (1999) and Müller et al. (2006) computed heliospheric models for a wide range of interstellar pressures, showing that when the heliosphere enters a cold cloud with density 100,000 times larger than the Local Interstellar Cloud (LIC) in which the heliosphere is now embedded, the heliosphere would shrink to the size of the inner solar system. The density and ionization within the heliosphere will respond to time variations in the interstellar density, ionization, flow vector, and magnetic field strength. In order to estimate the possible range of physical parameters in the heliosphere over time, it is essential to explore the range of interstellar environments that the heliosphere has and will encounter. Our understanding of heliospheric evolution provides the basis for modeling astrospheres and their interactions with exoplanet atmospheres.

In theoretical models of the ISM (e.g., Field et al. 1969; Wolfire et al. 1995, 2003), the balance between heating and cooling processes in a constant-pressure environment predicts that thermal instability will drive the ISM into distinct phases each characterized by a narrow range of temperatures and densities. These phases have been called the cold neutral medium (CNM; T < 300 K), warm neutral medium (WNM; T ≈ 8000 K), warm ionized medium (WIM; T ≈ 8000 K), and hot ionized medium (HIM; T ≈ 106 K). These theoretical models were constructed assuming steady-state populations and pressure equilibrium among the phases. In a more recent theoretical study that included revised cooling terms, Wolfire et al. (2003) argued that at the location of the Sun in the Galaxy, turbulent heating processes should not invalidate the pressure-equilibrium assumption. A very different model was found from simulations that include supernova and stellar wind heating and dynamics and time-dependent evolution of gas without the constraint of constant pressure. In the simulations of de Avillez & Breitschwerdt (2005), turbulence creates a wide range of temperatures, densities, magnetic field strengths, and flows. This model is consistent with 21 cm radio studies (Heiles & Troland 2003; Begum et al. 2010) that showed that a large fraction of the warm neutral gas has temperatures in the 500–5000 K range that are thermally unstable in the Wolfire et al. (2003) theoretical models of the WNM. Given these two very different models for the ISM, it is important to measure the properties of nearby interstellar gas because proximity provides the best angular resolution, and short path lengths to nearby stars minimizes the complexity inherent in long sight lines through the ISM. A critical question is whether the local interstellar medium (LISM) is more like the static or dynamic models, or perhaps contains aspects of both models.

This paper investigates the inhomogeneity of the LISM: the range of temperatures and turbulent velocity variations inside the LIC and the nearby partially ionized clouds in the Cluster of Interstellar Clouds (CLIC), possible trends in these inhomogeneities, possible causes for these inhomogeneities, and whether the LISM consists of warm clouds that partially or completely fill the LISM. A review paper, entitled "Inhomogeneity in the local ISM and its relation to the heliosphere" (Linsky et al. 2022), contains a preliminary presentation and discussion of these issues based upon a smaller data set. Other previous reviews include Redfield (2006) and Frisch et al. (2011).

In Section 2 we describe our measurement techniques and previous measurements of the temperatures, turbulent velocities and kinematics of nearby warm interstellar clouds. Section 3 describes the inhomogeneous properties within the LIC and other clouds, and possible trends with different parameters. In Section 4 we test the predictions of theoretical models against our data, and Section 5 summarizes our results. The Appendix provides a summary of the interstellar parameters in the LISM for all sight lines out to 100 pc.

2. Measurement of Cloud Temperatures and Turbulence

Our knowledge of the LISM is based primarily on high-resolution UV spectra that include narrow absorption lines produced by interstellar gas in the line of sight to a star. The most useful interstellar lines are transitions from the ground states of H i (Lyα, 1215.67 Å), D i (Lyα, 1215.34 Å), Mg ii (2796.35, 2803.53 Å), and Fe ii (2586.65, 2600.127 Å). Spectral lines of N i, O i, C ii, and Si ii require more deeply exposed spectra than most spectra analyzed in this paper to provide high signal-to-noise ratio interstellar absorption lines. Interstellar absorption lines of Ca ii and Na i in the optical spectrum are generally too weak to be seen on short sight lines.

The procedure for analyzing Hubble Space Telescope (HST) absorption line profiles is described in detail by Redfield & Linsky (2004a, 2004b) and Malamut et al. (2014). Here we provide a short overview. The high-resolution Space Telescope Imaging Spectrograph (STIS) spectra were obtained from the .x1d files in the MAST Portal. We used a suite of software written in the IDL language to fit Voigt profiles to the narrow interstellar absorption lines observed against broad emission lines formed in stellar chromospheres. We first fit the stellar emission line, typically with a self-reversal, by a sixth-degree polynomial, and then simultaneously fit one or more interstellar profiles centered at different velocities to obtain an optimal fit to the observed spectrum. For Mg ii we fit both lines in the multiplet simultaneously. We also individually fit both fine-structure components of the H i and D i Lyα lines. The fitting procedure takes into account instrumental broadening appropriate for the time of observation. The output consists of the central wavelength and velocity for each interstellar absorption line, the line-width Doppler parameter b, the column density of the atom or ion, the estimated hydrogen column density, and both formal errors and Monte Carlo simulation errors. For every sight line that we have analyzed, there is at least one interstellar absorption component; but there are often two or more velocity components, indicating different parcels of interstellar gas moving at different radial velocities in the sight line to the star. The average number of components is about 1.5 per sight line (Malamut et al. 2014).

The measurement of cloud temperatures T and nonthermal (turbulent) broadening parameter ξ involves comparing the line widths of multiple ions of different atomic mass. The interstellar absorption line width b is the sum of the thermal and turbulent components,

Equation (1)

where k is Boltzmann's constant, m is the mass of the atom or ion, and A is the atomic weight of the atom or ion in atomic mass units. This equation assumes that the observed plasma can be described by a single Maxwell–Boltzmann velocity distribution with macroscopic random flows along the line of sight through the interstellar cloud. This simple representation need not be the case if the temperature is not uniform along the line of sight or the plasma has a microscopic suprathermal component (see, Quémerais & Izmodenov 2002; Schwadron et al. 2011) that could produce extended wings or Doppler shifts in the line profile. While these complex phenomena are observed in the solar corona, they have not yet been investigated for interstellar clouds. We therefore assume that T and ξ represent mean values through a cloud and proceed with the standard formula until evidence emerges that requires a more detained analysis.

The separation of thermal from nonthermal macroscopic broadening follows from the different atomic weights of the observed interstellar atoms and ions. In principle, the separation only requires line widths obtained from high-resolution spectra of a low-mass atom such as D i and a high-mass ion such as Mg ii or Fe ii, because the line widths of a low-mass atom are primarily thermal, whereas the line widths of the high-mass ion are primarily due to gas motions along the sight line. The D i Lyα line separated from the H i Lyα line by −0.33 Å is preferred for interstellar studies, because it is generally optically thin whereas the H i line is very optically thick. The D i line is actually a hyperfine structure doublet with components at 1215.3430 and 1215.3373 Å. Because of the relatively weak thermal broadening of heavy ions, especially Fe ii, these ions best display the multiplicity of velocity components. However, velocity components separated by less than the resolution of the spectrograph are unresolved and a one-component analysis results in a large value for the nonthermal broadening. Thus high-resolution spectra are essential for this work.

Since the line widths range from b ≈ 3 km s−1 for Fe ii to b ≈ 8 km s−1 for D i, high-resolution spectra with an accurate absolute wavelength scale are essential. For the 1200–1700 Å spectral range, the best-available spectra are obtained with the E140H grating (3 km s−1 resolution) or E140M grating (6 km s−1) in the STIS instrument or the ECH-A (3.6 km s−1) grating in the earlier High Resolution Spectrograph (HRS) instrument. For the 2000–3000 Å spectral range the best-available spectra are obtained with the STIS E230H (3 km s−1) or HRS ECH-B (3.3 km s−1) gratings. Unless otherwise noted, all Mg ii and Fe ii line widths are derived from E230H spectra, and most H i, D i, C ii, and O i line widths from E140M spectra. We identify in Figure 1 the data obtained with E140H by red points and the data obtained with E140M with blue points. In the new spectra that we analyze, the C ii and O i lines are usually too faint to use.

Figure 1.
Standard image High-resolution image
Figure 1.

Figure 1. Left: plots of turbulent velocity vs. temperature for different ions, best fit (cross), and 1σ and 2σ contours about the best fit. Right: Doppler parameters vs. mass for each ion. The best-fit parameters and 1σ contours fitting the Doppler parameters are included. Data from E140M spectra are blue and data from E230H or E140H spectra are red.

Standard image High-resolution image

Figure 1 shows that the analysis of line widths of low- and high-mass ions provides the best estimate of the mean temperature and turbulent velocity in the sight lines to stars that we have measured.

The measured neutral hydrogen column density plays a critical role in determining the size of interstellar clouds (see Section 3.11), but the analysis of the only available neutral hydrogen line with interstellar absorption, Lyα, poses particular challenges: the interstellar line core is extremely optically thick, often 106 even for short sight lines, the line has broad damping wings that require a Voigt profile for fitting, there may be a central self-reversal produced by the loss of photons near the top of the chromosphere (see Youngblood et al. 2022), and the narrow geocoronal emission feature located near line center must be removed before fitting the line profile. In addition, charge exchange between protons and hydrogen atoms flowing into the heliosphere or astrosphere from the LISM produces regions of decelerated, hot, and relatively dense neutral hydrogen in what are called hydrogen walls in the outer heliosphere and astrospheres of stars similar to and cooler than the Sun. Wood et al. (2005a) described their analysis of Lyα lines that take into account the extra absorption on the red side of the interstellar absorption from the heliospheric hydrogen wall and the extra absorption mostly on the blue side from the astrospheric hydrogen wall. Since these extra absorptions can be weak and difficult to identify and fit accurately, the analysis of the hydrogen Lyα line by itself needs a further constraint. The neutral deuterium Lyα line centered at −0.33 Å from the hydrogen Lyα line provides this constraint, because the deuterium line optical depth is often optically thin and N(D i) is too small for any significant deuterium wall absorption. The neutral hydrogen column density can be inferred from the deuterium line, N(H i) = N(D i)/R, where R = 1.5 × 10−5 is the deuterium/hydrogen number ratio within 100 pc of the Sun (Wood et al. 2004; Linsky et al. 2006). Most N(H i) data comes from simultaneously fitting both the hydrogen and deuterium lines, but highly constrained N(H i) results can be obtained by fitting only the deuterium line.

2.1. Previous Measurements

The first measurements of interstellar temperatures, turbulent velocities, and kinematics were for individual bright stars, e.g., Capella (Linsky et al. 1993), Sirius A (Lallement et al. 1994), Procyon (Linsky et al. 1995), epsilon CMa (Gry et al. 1995; Gry & Jenkins 2001), α Cen (Linsky & Wood 1996), and β CMa (Jenkins et al. 2000). Redfield & Linsky (2004a) and Redfield & Linsky (2008) provided the first analysis of a large number of sight lines using the spectrographs on board HST. They measured temperatures, nonthermal broadening, and heliospheric velocities for 50 velocity components in the sight lines toward 29 stars located within 100 pc. Fourteen of the sight lines show only one velocity component, nine show two components, and six show three components. All of the stars show at least one interstellar-velocity component. Thus interstellar absorption observed in the H i, D i, and Mg ii lines is ubiquitous with multiple velocity components common even for lines of sight as short as 2.6 pc.

Redfield & Linsky (2004b) found a weighted mean temperature and weighted dispersion $ \langle T \rangle ={6680}_{-}^{+}1490$ K. The distribution does not appear Gaussian about the mean value as there are at least eight velocity components with temperatures well below that expected for a Gaussian distribution with this dispersion. Additional observations are needed to better characterize the distribution of temperatures. They also found that the mean value of nonthermal broadening is 〈ξ〉 = 2.24 ±1.03 km s−1 with an apparent excess of large ξ values that could result from unresolved closely spaced velocity components. Subsequently, Redfield & Linsky (2008) found a weighted mean temperature of 7500 ± 1300 K for the 19 LIC stars in their sample. Table 1 compares the weighted mean temperatures and weighted dispersions about the mean temperatures obtained from past and present data sets. For comparison, we include the temperature of neutral helium flowing into the heliosphere from the LIC including corrections for heating and deceleration resulting from elastic collisions in the VLISM (Swaczyna et al. 2021). There is excellent agreement between the new mean temperatures for the LISM and LIC data sets with the temperatures measured for the inflowing neutral helium.

Table 1. Temperature Measurements for Local Interstellar Gas

RegionComponentsMean TemperatureTemperature RangeReference
Inflow of LIC He i gas16400 KSwaczyna et al. (2021)
LIC197500 ± 1300 K5200–12,900 KRedfield & Linsky (2008)
LIC (all data)376509 ± 2761 K1830–12,900 KThis paper
LIC (only 2σ data)366511 ± 2773 K2450–12,900 KThis paper
CLIC506680 ± 1490 K1000–12,600 KRedfield & Linsky (2004b)
CLIC (all data) 97 6742 ± 2,441 K1700–12,900 KThis paper
CLIC (only 2σ data)846838 ± 2455 K2450–12,900 KThis paper

Download table as:  ASCIITypeset image

Following the Redfield & Linsky (2008) paper there have been two papers with additional values of T and ξ for other sight lines. Zachary et al. (2018) studied two sight lines each with two velocity components, and Edelman et al. (2019) studied three stars each with three velocity components. The results of these two studies are included in Table 2. The results of our analysis of 27 new sight lines with 31 velocity components are listed in Table 3. In the last column of this table, we list the likely cloud in the sight line; clouds are shown in parenthesis when the sight line passes through its edge. The addition of these 47 velocity components listed in Tables 2 and 3 to the initial list of 50 sight lines obtained by Redfield & Linsky (2004b) warrants a reexamination of the distribution of cloud temperatures and turbulent velocities. The analysis of the new sight lines in this paper followed the approach and used the same software as described in detail by Redfield & Linsky (2004a).

Table 2. Temperature and Turbulent Velocity Measurements since 2008

HDName l b d (pc)v T ξ Ions UsedRef
209100 epsilon Ind336.2−48.03.62−40.4 ${8340}_{-440}^{+450}$ ${1.97}_{-0.09}^{+0.08}$ H i, Mg ii, Fe ii 4
190248 δ Pav329.8−32.46.10−17.2 ± 1.5 ${8680}_{-780}^{+740}$ ${0.0}_{-0.0}^{+2.18}$ D i, C ii, O i, Mg ii 1
190248 δ Pav329.8−32.46.10−9.23 ± 0.58 ${9310}_{-7959}^{+10070}$ ${2.44}_{-2.44}^{+1.04}$ D i 1
192310GJ 78515.6−29.48.81−30.41 ± 0.57 ${8600}_{-1800}^{+2000}$ ${3.3}_{-1.3}^{+1.2}$ D i, C ii, Mg ii, Fe ii 2
192310GJ 78515.6−29.48.81−24.24 ± 0.50 ${9900}_{-2100}^{+2200}$ ${2.2}_{-2.0}^{+1.1}$ D i, C ii, Mg ii, Fe ii 2
192310GJ 78515.6−29.48.81−18.88 ± 0.57 ${6900}_{-2300}^{+2600}$ ${1.3}_{-1.3}^{+1.6}$ D i, C ii, Mg ii, Fe ii 2
HIP85665GJ 678.1A28.620.510.12−31.5 ± 2.4 $10,{720}_{-3860}^{+4730}$ ${0.0}_{-0.0}^{+1.92}$ D i, Mg ii 1
HIP85665GJ 678.1A28.620.510.12−23.9 ± 2.4 ${8540}_{-2790}^{+3240}$ ${0.0}_{-0.0}^{+1.34}$ D i, Mg ii 1
9826 υ And132.0−20.713.489.1 ± 1.2 $10,{400}_{-1900}^{+2000}$ ${0.00}_{-0.0}^{+2.4}$ D i, O i, Mg ii, Fe ii 2
9826 υ And132.0−20.713.4812.1 ± 1.1 ${4000}_{2200}^{+2800}$ ${1.8}_{-1.2}^{+0.8}$ D i, O i, Mg ii, Fe ii 2
9826 υ And132.0−20.713.4816.45 ± 0.88 ${6500}_{-2600}^{+3000}$ ${1.3}_{-1.3}^{+0.7}$ D i, O i, Mg ii, Fe ii 2
206860NN Peg69.9−28.318.13−14.68 ± 0.58 ${7100}_{-2400}^{+2800}$ ${1.4}_{-1.4}^{+0.6}$ D i, Mg ii, Fe ii 2
206860NN Peg69.9−28.318.13−8.0 ± 1.0 ${9600}_{-2300}^{+2500}$ ${2.11}_{-0.68}^{+0.54}$ D i, Mg ii, Fe ii 2
206860NN Peg69.9−28.318.13−5.44 ± 0.79 ${6800}_{-2500}^{+2700}$ ${0.80}_{-0.80}^{+0.98}$ D i, Mg ii, Fe ii 2
87901 α Leo226.448.924.318.8 ± 0.2 ${6000}_{-600}^{+600}$ 1.78 ± 0.10C ii, N i, O i, Mg i, Mg ii 3
87901 α Leo226.448.924.3114.4 ± 0.1 ${5990}_{-700}^{+700}$ 1.85 ± 0.19C ii, N I, O i, Mg I, Mg ii 3

References. (1) Zachary et al. (2018); (2) Edelman et al. (2019); (3) Gry & Jenkins (2017); Malamut et al. (2014).

Download table as:  ASCIITypeset image

Table 3. New Temperature and Turbulent Velocity Measurements

HDName l b d (pc)v T ξ Ions UsedRefCloud
217987GJ 8875.1−66.03.29−2.74 ± 0.21 ${6000}_{-1400}^{+1500}$ ${1.65}_{-1.10}^{+0.72}$ D i, Mg ii 1,2LIC
1326GJ 15A116.7−18.43.5610.91 ± 0.32 ${7800}_{-1100}^{+1200}$ ${2.81}_{-0.21}^{+0.20}$ D i, Mg ii 1, 2(Hya)
 GJ 273212.310.43.8018.28 ± 0.928620 ± 350 ${0.0}_{-0.0}^{+1.32}$ H i, Mg ii 1,2LIC
 GJ 273212.310.43.8021.38 ± 0.40 ${8620}_{-340}^{+350}$ ${0.0}_{-0.0}^{+1.59}$ H i, Mg ii 1,2(Aur)
239960AGJ 860A104.7−0.04.01−0.14 ± 0.67 ${4780}_{-1780}^{+2080}$ ${2.24}_{-0.51}^{+0.42}$ D i, Mg ii 1(Eri)
GJ 873EV Lac100.6−13.15.04.44 ± 0.43 ${3030}_{-610}^{+670}$ ${2.21}_{-0.23}^{+0.22}$ D i, Mg ii 1(LIC)
36395GJ 205206.9–19.45.7017.31 ± 0.19 ${4200}_{-680}^{+680}$ ${2.37}_{-0.24}^{+0.24}$ H i, D i, Mg ii 1, 2unassigned
36395GJ 205206.9–19.45.7021.93 ± 0.16 ${4500}_{-1200}^{+1200}$ ${3.64}_{-0.27}^{+0.27}$ H i, D i, Mg ii 1, 2(LIC)
 GJ 588332.712.15.92−26.58 ± 0.56 ${6730}_{-720}^{+700}$ ${3.37}_{-0.31}^{+0.32}$ H i, Mg ii 1,2G
 YZ CMi215.913.55.9918.11 ± 0.318930 ± 260 ${0.0}_{-0.0}^{+1.18}$ H i, Mg ii, Fe ii 1,2LIC
 YZ CMi215.913.55.9921.73 ± 0.538790 ± 230 ${1.59}_{-0.77}^{+0.57}$ H i, Mg ii, Fe ii 1,2(Aur)
191408GJ 7835.2−30.96.01−25.9 ± 0.42 ${7700}_{-1500}^{+1500}$ ${3.05}_{-0.86}^{+0.86}$ D i, Mg ii 1Mic
79210GJ 338A164.942.76.3312.58 ± 0.09 ${6650}_{-1020}^{+970}$ ${2.29}_{-0.55}^{+0.61}$ D i, Mg ii, Fe ii 1,2LIC
11561761 Vir.311.944.18.57−16.7 ± 0.47 ${6600}_{-4400}^{+4800}$ ${1.60}_{-1.60}^{+0.83}$ D i, C ii, O i 1(NGP)
23249 δ Eri198.1−46.09.0420.29 ± 0.43 ${3650}_{-1110}^{+1200}$ ${2.25}_{-0.22}^{+0.19}$ D i, C ii, Mg ii 1LIC
37394GJ 211158.411.912.317.2 ± 0.27 ${5600}_{-2500}^{+2200}$ ${1.00}_{-1.00}^{+0.71}$ D i, C ii, Mg ii 1LIC
166HR 8111.3−32.813.76.50 ± 0.44 ${4520}_{-1880}^{+2110}$ ${4.39}_{-0.34}^{+0.33}$ D i, C ii, Mg ii 1LIC
72905 π1 UMa150.635.714.512.91 ± 0.43 ${2450}_{-660}^{+740}$ ${2.47}_{-0.11}^{+0.10}$ D i, Mg ii 1LIC
142373 χ Her67.750.315.8−12.69 ± 0.16 ${4700}_{-1100}^{+1200}$ ${1.36}_{-0.41}^{+0.31}$ D i, Mg ii, Fe ii 1(Mic)
43162GJ 3389230.9−18.516.717.1 ± 0.43 ${3200}_{-2330}^{+2680}$ ${5.76}_{-0.26}^{+0.25}$ D i, Mg ii 1LIC
165185GJ 702.1356.0−07.317.2−26.9 ± 0.43 ${8610}_{-3340}^{+4150}$ ${0.0}_{-0.00}^{+1.76}$ D i, Mg ii 1(G)
116956SAO 28753113.759.521.7−0.25 ${5160}_{-2680}^{+3420}$ ${1.79}_{-1.37}^{+0.69}$ D i, Mg ii 1(LIC)
116956SAO 28753113.759.521.77.7 ${2540}_{-1830}^{+2450}$ ${1.66}_{-1.66}^{+1.21}$ D i, Mg ii 1unassigned
59967GJ 3446250.5−09.021.89.32 ± 0.02 ${12400}_{-2700}^{+2900}$ ${0.00}_{-0.0}^{+1.9}$ D i, C ii, O i 1Blue
97334GJ 417184.367.322.63.34 ± 0.47 ${11000}_{-3000}^{+2400}$ ${0.0}_{-0.0}^{+1.4}$ D i, C ii, O i 1LIC
73350.GJ 9273232.120.024.311 ${7900}_{-3200}^{+3200}$ ${0.68}_{-0.68}^{+1.67}$ D i, Mg ii 1LIC
82210DK UMa142.538.932.49.41 ± 0.61 ${4220}_{-430}^{+460}$ ${1.86}_{-0.18}^{+0.15}$ D i, Mg ii 1(LIC)
129333EK Dra105.549.034.6−2.43 ± 0.32 ${2580}_{-250}^{+280}$ ${3.13}_{-0.05}^{+0.05}$ D i, C ii, Mg ii 1 (LIC)
93497 μ Vel283.008.635.9−4.38 ± 0.43 ${3470}_{-590}^{+580}$ ${3.12}_{-0.10}^{+0.11}$ D i, Mg ii 1G
128987KU Lib331.030.242.3−20.90 ± 0.62 ${6580}_{-3680}^{+4500}$ ${3.00}_{-0.10}^{+0.11}$ D i, Mg ii 1Gem
209458V376 Peg76.8−28.548.1−29.17 ± 0.94 ${8500}_{-2850}^{+3280}$ ${0.00}_{-0.00}^{+3.02}$ D i, Mg ii 1unassigned

References. (1) This paper; (2) Wood et al. (2021).

Download table as:  ASCIITypeset image

3. Analyzing the Distributions of Cloud Properties

Until recently, interstellar clouds in the LISM have been characterized by the mean properties measured for all sight lines passing though the cloud consistent with the cloud's velocity vector. With the availability of 97 velocity components passing through the LISM and 37 of these passing though the LIC, we can begin to do statistical analyses of the data. In the following analyses, we include only those velocity components for which the temperatures are at least two times larger than the measurement errors, because the software may not fully separate thermal from turbulent velocity, and there could be unresolved velocity components that would appear as extra broadening. For the entire data set 84 of the 97 velocity components meet this 2σ criterion, and for velocity components passing through the LIC 36 of the 37 meet this criterion. In this analysis we include sight lines out to stars only within 100 pc. We searched for trends in physical properties by asking specific questions.

3.1. Are the Temperatures and Turbulent Velocities within a Cloud Roughly Constant or Variable?

An important question is whether the cloud temperatures and turbulent velocities within a cloud are roughly constant or distributed in a random or systematic manner. If these properties are random, then their distributions could be Gaussian distributed with perhaps a few outliers. With the 84 sight lines now evaluated for temperature and turbulent velocity, we can address this question. Figure 2 (left) shows the number of temperature measurements in each 1000 K temperature bin. This bin size is appropriate given the typical measurement uncertainties of 500–2000 K. The weighted mean gas temperature is 6838 ± 2455 K, and the solid line in the plot is a Gaussian fit to the data. The fit is reasonably good, although there are seven outliers in the temperature range 12,000–13,000 K. The errors are weighted dispersions about the mean (see Redfield & Linsky 2004b). Table 1 lists the weighted mean temperatures and dispersions measured in different ways. Figure 2 (right) shows the distribution of temperatures for the 36 sight lines that pass through the LIC. While there are fewer sight lines in this plot, the distribution, mean value, and dispersion of the temperatures are similar to the plot that included all sight lines. These data show that the LIC and the other warm clouds in the CLIC have similar wide distributions of internal temperatures.

Figure 2.

Figure 2. Distributions of sight-line temperatures for the full data set (left) and for only the sight lines that pass through the LIC (right). Temperatures are binned in 1000 K intervals. The solid curves are Gaussian fits to the weighted data points.

Standard image High-resolution image

Figure 3 (left) shows the distribution of turbulent velocities plotted with a bin size of 0.5 km s−1 as typical measurement uncertainties are in the range 0.2–1.5 km s−1. The weighted mean turbulence is 2.54 ± 1.18 km s−1, and the solid line Gaussian in the figure is also a good fit to the data. The distribution of turbulent velocities for the sight lines that pass through the LIC (Figure 3, right) has similar properties. We conclude that wide ranges of temperatures and turbulent velocities characterize the gas in the nearby warm clouds. As previously noted by Redfield & Linsky (2004a), the turbulent velocities are subsonic and the highest turbulent velocities may result from unresolved velocity components.

Figure 3.

Figure 3. Distributions of sight-line turbulent velocities for the full data set (left) and for only the sight lines that pass through the LIC (right). Turbulent velocities are binned in 0.5 km s−1 intervals. The solid curves are Gaussian fits to the weighted data points.

Standard image High-resolution image

Another method for testing whether the temperatures and turbulent velocities within the LIC are constant or have variable properties is to compare the properties of pairs of sight lines as a function of their angular separation. For the 36 sight lines traversing the LIC there are 36 × 35 = 1260 pairs, but only 596 unique pairs after subtraction of pairs for two clouds in the sight line to the same star and pairs that are sampled twice by the search software. Figure 4 (left) shows the temperature differences for these pairs as a function of angular separation between the sight lines. The mean value of these temperature differences is 2845 K, but there are many temperature differences exceeding 6000 K. Since the mean value of the individual measurement uncertainties for these sight lines is 1625 K, most of these sight-line pairs have temperature differences that well exceed the measurement errors.

Figure 4.

Figure 4. Left: temperature differences for sight-line pairs through the LIC as a function of angular separation in degrees between the sight lines. The horizontal red line is the mean uncertainty of temperature measurements. Right: mean temperature differences for sight-line pairs through the LIC in 10° angular bins. The vertical lines show the variance of the temperature differences in each angular bin and the star symbols are the mean value in each bin. The blue line is a least-squares fit to the mean temperature differences as a function of angular separation.

Standard image High-resolution image

Figure 4 (right) plots the same data in angular separation bins 10° wide. The blue line is a least-squares fit to the mean values in each bin. The line slope is 1.96 times its error, indicating a significant increase in the mean temperature differences between zero angular separation (2465 K) and those at the widest separation. This result indicates significant temperature differences on all angular separation scales, even at the smallest separations.

3.2. What Is the Length Scale for the Temperature Inhomogeneities in the LIC?

There are many studies of inhomogeneous properties in the ISM; see, for example, the comprehensive review by Stanimirović & Zweibel (2018). Very-small-scale structures of neutral cold gas are inferred from high-resolution absorption lines of neutral carbon (Jenkins & Tripp 2007), sodium (Lauroesch 2007), hydrogen (Stanimirović et al. 2007), and molecules with tiny-scale atomic structures as small as tens of au. Very small structures of ionized gas in the diffuse ISM also show tiny structures with sub-au scales. These very-small-scale structures are identified at radio wavelengths and are caused by variations in the electron density and therefore the refractive index along the sight line that result in diffraction patterns called scintillation; see the reviews by Rickett (2007) and Cordes et al. (2007). Structure sizes as small as 1 au are inferred from the rotation measure structure functions and from scintillation arcs produced by highly turbulent density clumps, called scattering screens, along sight lines to pulsars and quasars. Linsky et al. (2008) found three scattering screens within a few parsecs of the Sun where the edges of two or more partially ionized clouds may interact. These small spatial scales for both neutral and ionized gas refer to density and likely magnetic field fluctuations, and are based on measurements of neutral and ionized gas through long sight lines. With the available data, we can now measure the length scales for temperature fluctuations in the immediate environment of the Sun, including the LIC and other partially ionized clouds, and what these fluctuations suggest could be the timescale of externally driven changes in the heliosphere. Are the length scales for temperature fluctuations in local partially ionized warm clouds similar to the length scales for electron density fluctuations in the ionized ISM?

The least-squares fit in Figure 4 (right) predicts that the mean temperature difference for sight lines with the smallest angular separations should be about 2465 K, but the mean temperature measurement error is only 1625 K. For stars in our data set, the smallest angular scale of 2fdg2 is for the Procyon–YZ CMi sight-line pair, which has a temperature difference of 2220 K, but the measurement uncertainty for this pair is only 695 K, a factor of 3.2 smaller. Table 4 lists the four sight-line pairs with the closest angular separations. We note that the H i column densities are the same for the first three pairs, consistent with the shape of the LIC not being highly irregular. The larger and more discordant log N(H i) values for the fourth pair may be due to additional column densities through the edges of the Eri and Hyades clouds. We estimate the path length d(LIC) through the LIC from the neutral hydrogen column densities and n(H i) = 0.20 cm−3. The separation, s, of the sight lines halfway through the LIC is then s = 206265 × (d(LIC)/2)tan(θ) au, where θ is the angular separation of the sight lines. The separations range from 5100 au to 17,360 au for these close pairs. These separations could be upper limits to the true inhomogeneity length scale, because there could be unmeasurable but significant temperature variations within the sight lines to even the closest stars and between the stars with the smallest angular separations. Since the Sun moves through the LIC at 5.1 au yr−1, the heliosphere could see changes in local interstellar properties within 1000 yr. We conclude that the linear scale for significant temperature differences in the LIC is at least as small as 5100 au.

Table 4. Sight-line Separations Halfway Through the LIC for the Closest Pairs

AngleStar 1Star 2ΔT UncertaintyΔT/Uncertaintylog N(H i) d (LIC)Separation
2fdg20ProcyonYZ CMi2220 K695 K3.2 17.9, 17.891.29 pc5100 au
2fdg62 epsilon Eri δ Eri3760 K1435 K2.617.88, 17.881.29 pc6,075 au
2fdg94ProcyonGJ 2731910 K733 K2.617.9, 17.861.23 pc6,510 au
3fdg12HD 166PW And6780 K2724 K2.518.3 a , 18.1 a 3.09 pc17,360 au

Note.

a A portion of log N(H i) for HD 166 may come from the Eri and Hyades clouds.

Download table as:  ASCIITypeset image

This technique for estimating length scales is similar to that described by Spangler (2001), who compared rotation measures as a function of angle between different sight lines. Such measures are proportional to the integral of the product of the electron density and the projection of a magnetic field along a line of sight. There is a break in the difference between rotation measures for sight lines with angular separations less than or about 0fdg1, corresponding to a length scale for these plasma fluctuations of about 3.6 pc (Minter & Spangler 1996).

3.3. Are Temperatures and Turbulent Velocities Correlated?

Figure 5 (left) plots temperature versus turbulent velocity for all 84 velocity components. Although there is a large scatter in the data, there is a clear trend of decreasing temperature with increasing turbulent velocity. For the full data set, the least-squares linear fit is in the form T = A + Bvturb, where A = 8816 ± 449 K and B = −795 ± 220, with the slope 3.6 times its error. For the LIC-only data (Figure 5 (right)), the fit is A = 9350 ± 523 with a steeper slope B = −1523 ± 287 that is 5.3 times its error.

Figure 5.

Figure 5. Plots of sight-line temperatures vs. turbulent velocities for the full data set (left) and the LIC-only data set (right). The solid lines are least-squares linear fits to the data.

Standard image High-resolution image

With the smaller data set then available, Redfield & Linsky (2004b) found a moderately negative correlation between temperature and turbulent velocity with a Pearson r = −0.47 for the entire sample and r = −0.35 for the subsample with more precise parameters. For the larger sample shown in Figure 5 (left) the correlation coefficient is r = −0.37, and for the LIC-only data set the coefficient is larger, r = −0.67, consistent with visual inspection. The covariance for the two data sets are −1077 and −1842, respectively.

This significant correlation of temperature and turbulent velocity, especially in the new LIC data set, could result in part from the absorption-line-fitting procedure, because the fit to low-mass atoms (e.g., D i) and high-mass ions (e.g., Mg ii and Fe ii) depends on the sum of thermal and turbulent broadening. For example, a positive (or negative) error in the temperature measurement can be partially compensated by a negative (or positive) error in the turbulent velocity measurement. The presence of unresolved velocity components would significantly increase the measured turbulent velocities with only a small effect on the temperature measurements. Since the effect of unresolved velocity components is to produce the largest turbulent velocities, the correction of these data points would produce an even steeper correlation. Redfield & Linsky (2004b) discuss other possible causes of the correlation including systematic errors. If the correlation is real and not an artifact of the measurement technique or inadequate spectral resolution, then it could indicate the conversion of turbulent energy to heat. Table 5 summarizes the A and B coefficients and Pearson r values for all of the linear fits. The underlined values of r indicate significant trends.

Table 5. Linear Least-squares Fits for y = A + Bx

Plot for parameters y versus x Data set (2σ) A B Pearson r
Temperature versus turbulent velocityAll8816 ± 449−795 ± 220 –0.371
Temperature versus turbulent velocityLIC9350 ± 523−1523 ± 287 –0.672
Temperature versus distanceAll6969 ± 37928.7 ± 15.0+0.207
Temperature versus angle from Galactic CenterAll6804 ± 6147.20 ± 5.86+0.135
Temperature versus angle from epsilon CMaAll7704 ± 6272.53 ± 6.42−0.044
Temperature versus angle from epsilon CMaLIC7342 ± 1010−2.83 ± 11.90−0.041
Temperature versus angle from inflow directionAll6669 ± 6358.42 ± 5.96+0.154
Temperature versus angle from LIC coreLIC6127 ± 103217.01 ± 16.16+0.178
Temperature versus angle from magnetic fieldAll6352 ± 74111.39 ± 6.97+0.178
Temperature versus angle from magnetic fieldLIC7298 ± 1530−1.549 ± 13.031−0.020
Temperature versus N(H i)All7657 ± 13068−9.81 ± 731.1−0.001
Temperature versus N(H i)LIC171434 ± 19113−560 ± 1067−0.090
Temperature difference versus angle from LIC coreLIC2465 ± 2364.72 ± 2.40 +0.452
Turbulent velocity versus distanceAll1.764 ± 0.180−0.00468 ± 0.00713−0.072
Turbulent velocity versus angle from Galactic CenterAll1.612 ± 0.2890.000722 ± 0.00276+0.029
Turbulent velocity versus angle from epsilon CMaAll1.952 ± 0.291−0.00310 ± 0.00298−0.114
Turbulent velocity versus angle from epsilon CMaLIC1.239 ± 0.4440.00289 ± 0.00524+0.094
Turbulent velocity versus angle from LIC coreLIC1.596 ± 0.463−0.00228 ± 0.00724−0.054
Turbulent velocity versus angle from inflow directionAll1.623 ± 0.3000.000592 ± 0.00281+0.023
Turbulent velocity versus angle from magnetic fieldAll1.661 ± 0.3510.00019 ± 0.00330+0.006
Turbulent velocity versus angle from magnetic fieldLIC1.157 ± 0.6730.00271 ± 0.00574+0.081
Turbulent velocity versus angle from magnetic field (bins)All1.250 ± 0.4620.00385 ± 0.00445+0.311
Turbulent velocity versus angle from magnetic field (bins)LIC0.496 ± 0.7270.00993 ± 0.00701+0.472
Turbulent velocity versus N(H i)All−4.33 ± 6.200.335 ± 0.347+0.106
Turbulent velocity versus N(H i)LIC−4.69 ± 8.410.339 ± 0.470+0.123
Mean density versus thermal pressure160.0338 ± 0.0312(4.48 ± 4.79) × 10−6 +0.243
Mean density versus thermal+turbulent pressure160.0336 ± 0.0291(2.84 ± 2.81) × 10−10 +0.261

Download table as:  ASCIITypeset image

3.4. Do Temperatures and Turbulent Velocities Depend on Distance to the Star?

We searched for evidence of whether the distance to the star at the end of the sight line influences the properties of the intervening interstellar gas. Figure 6 (left) plots temperature versus the distance to the star at the end of a given sight line. The absence of a significant trend in the data supports the assumption that background stars only serve as illumination sources. The fit parameter for the turbulent velocity data in Figure 6 (right) also shows no significant trend. A possible trend with distance to stars with very strong extreme-UV (EUV) emission is tested in Section 3.6. We can therefore combine sight lines of near and more distant stars when studying individual clouds.

Figure 6.

Figure 6. Plots of sight-line temperatures (left) and turbulent velocities (right) vs. distance to the star. The solid lines are least-squares linear fits to the data.

Standard image High-resolution image

3.5. Are the Temperatures and Turbulent Velocities of the LIC and G Clouds Different?

The heliosphere is now located near the edge of the LIC and will likely leave the LIC in less than 2000 yr (Redfield & Linsky 2004b; Frisch et al. 2011; Linsky et al. 2022). The trajectory is in the direction of the G Cloud, but whether the heliosphere will directly enter the G Cloud or an intercloud boundary consisting of Local Cavity gas or compressed partially ionized gas has been a matter of speculation. Very recently, Swaczyna et al. (2022) proposed that the heliosphere is now surrounded by overlapping LIC and G Cloud plasma with an average temperature intermediate between the two clouds and n(H i) = 0.20 cm−3, the sum of the densities of the two clouds. With our new data set, we can test whether the LIC and G Cloud have different properties. Table 6 lists the weighted mean values and weighted dispersions of the temperatures and turbulent velocities of the two clouds measured by Redfield & Linsky (2008) and measured in this paper. There are measured parameters for six sight lines through the G Cloud, but only four sight lines have temperatures that meet the 2σ criterion (α Cen A, α Cen B, 30 Oph, and μ Vel). As shown in Table 6, the weighted mean temperature is cooler and the weighted mean turbulent velocity is higher in the G Cloud than in the LIC, but the large weighted dispersions encompass the parameters for the two clouds.

Table 6. Mean Temperatures and Turbulent Velocities for the LIC and G Cloud

ParameterNumberLICNumberG CloudReference
Temperature197,500 ± 1,300 K55,500 ± 400Redfield & Linsky (2008)
Temperature366,511 ± 2,773 K45191 ± 847this paper
Turbulent velocity191.62 ± 0.7552.2 ± 1.1Redfield & Linsky (2008)
Turbulent velocity362.58 ± 1.3442.84 ± 1.05this paper

Download table as:  ASCIITypeset image

3.6. Do Temperatures and Turbulent Velocities Depend on Angle from the Inflow Direction?

Inflowing interstellar ions and ions produced by charge exchange reactions with inflowing interstellar neutrals are constrained by the heliospheric magnetic field to flow around the heliosphere away from the inflow direction. We therefore ask whether there are correlations of the measured temperatures and turbulent velocities with angle from the inflow direction. The upwind direction of interstellar gas flowing into the heliosphere is the vector sum of the interstellar gas flow in the local standard of rest and the Sun's motion relative to the local standard of rest. The upwind velocity, temperature, and direction have been studied using observations with the EUVE, IBEX, Ulysses, and STEREO spacecraft. A recent analysis of IBEX data by Swaczyna et al. (2021) gives T = 6400 K, v = 25.85 km s−1 and inflow Galactic coordinates (l = 3fdg5, b = +15fdg2) (Swaczyna et al. 2018). The temperatures and turbulent velocities in Figure 7 show no trend with angle θ from the upwind direction.

Figure 7.

Figure 7. Plots of temperatures (left) and turbulent velocities (right) vs. angle from the He i inflow direction (l = 3fdg5, b = 15fdg2).

Standard image High-resolution image

3.7. Do Temperatures and Turbulent Velocities Depend on the Angle from the Strong Ionizing Source epsilon CMa?

The strongest source of EUV radiation that ionizes hydrogen (λ < 912Å) is the star epsilon CMa (Galactic longitude l = 239fdg8 and Galactic latitude b = −11fdg3), but there are other stellar (Vallerga 1998) and hot white dwarf (Welsh et al. 2013) EUV sources distributed across the sky. Since β CMa, the second brightest EUV source, and Sirius B, the closest white dwarf, are in a similar direction as epsilon CMa, we searched to see whether there are correlations of temperature and turbulent velocity with angle from epsilon CMa. Figure 8 (left) shows the dependence of temperature and on angle from epsilon CMa for the full data set, and Figure 8 (right) shows the dependence for the LIC-only data set. We see no such effect for the temperature data or the turbulent velocity data (see Figure 9), with respect to the angle from epsilon CMa. The absence of any correlation with direction from epsilon CMa could be due to the many EUV sources distributed over the sky and the main heating source in the ISM being the photodissorption of electrons from dust grains that can be liberated by lower-energy photons.

Figure 8.

Figure 8. Plots of sight-line temperatures for the full data set (left) and the LIC-only data set (right) vs. angle from epsilon CMa. The solid lines are least-squares linear fits to the data.

Standard image High-resolution image
Figure 9.

Figure 9. Plots of sight-line turbulent velocities for the full data set (left) and the LIC-only data set (right) vs. angle from epsilon CMa. The solid lines are least-squares linear fits to the data.

Standard image High-resolution image

3.8. Do Turbulent Velocities Depend on Angle from the Magnetic Field Direction?

Spangler et al. (2011) suggested that turbulence in the local warm clouds could be anisotropic with velocity fluctuations that are larger perpendicular to the magnetic field direction than in the parallel direction. The basis for this suggestion was that in the solar wind and corona fluctuations in the magnetic field and plasma flow velocity are highly correlated and predominately perpendicular to the magnetic field direction. While this anisotropy decreases with heliocentric distance and likely decreases further as a result of ion-neutral collisions in the local clouds, Spangler et al. (2011) searched for but found no conclusive evidence for anisotropy in the data then available and encouraged a search in larger data sets as they become available. With the larger data set now available, we investigated the dependence of turbulent velocity on the angle from the magnetic field direction using the magnetic field direction l = 25fdg98, b = 50fdg09 in Galactic coordinates that Zirnstein et al. (2016) obtained from analysis of the IBEX ribbon data. They proposed that this direction is valid for the interstellar magnetic field at a distance of 1000 au.

Figure 10 shows the angular distribution of turbulent velocities relative to the magnetic field direction in 20° wide bins for the entire data set and for only those sight lines that traverse the LIC. In both plots there is no significant enhancement in the turbulent velocities near 90°, perpendicular to the magnetic field direction. A similar analysis of the temperatures also reveals no enhancement near 90°.

Figure 10.

Figure 10. Plots of sight-line turbulent velocities for the full data set (left) and the LIC-only data set (right) vs. angle from the magnetic field direction. The data are plotted in 20° bins. Asterix symbols indicate averages in each bin, and the vertical lines indicate the variance in the data. Asterix symbols with no vertical lines indicate only one datum in the bin.

Standard image High-resolution image

3.9. Do Temperatures and Turbulent Velocities Depend on whether Sight Lines Pass through the Core or Edge of the LIC?

Figure 11 addresses the question of whether the temperatures and turbulent velocities depend on whether sight lines penetrate through the core of the LIC (approximate coordinates l = 145° and b = 0°) or its edge. There are no significant trends in the temperature or turbulent velocity on direction of sight lines through the LIC, indicating that self-shielding from external EUV radiation sources is not important for determining the temperature or turbulent velocity of the LIC plasma.

Figure 11.

Figure 11. Plots of sight-line temperatures (left) and turbulent velocities (right) vs. angle from the center of the LIC. The solid lines are least-squares linear fits to the data.

Standard image High-resolution image

3.10. Do Temperatures and Turbulent Velocities Depend on the Neutral Hydrogen Column Density?

We show plots of the temperature (Figure 12) and turbulent velocity (Figure 13) versus N(H i) for the full data set and the LIC-only data set. There are no significant trends of either the temperatures or turbulent velocities on N(H i). One might expect higher temperatures in low-N(H i) sight lines that would be in closer proximity to the external EUV radiation field or perhaps hot gas surrounding the clouds, but there is no evidence that either scenario is the case.

Figure 12.

Figure 12. Plots of temperatures for the full data set (left) and the LIC-only data set (right) vs. the log of the neutral hydrogen column density. The solid lines are least-squares linear fits to the data.

Standard image High-resolution image
Figure 13.

Figure 13. Plots of turbulent velocities for the full data set (left) and the LIC-only data set (right) vs. the log of the neutral hydrogen column density. The solid lines are least-squares linear fits to the data.

Standard image High-resolution image

3.11. Do All Velocity Components Seen in the Mg ii and Fe ii lines Correspond to Warm Clouds?

Line-of-sight measurements of interstellar absorption provide information on the interstellar properties of gas in front of a star, but not where the cloud's absorbing gas begins and ends along the sight line. For stars within about 10 pc of the Sun, the detection of Lyα absorption blueshifted relative to the interstellar hydrogen absorption indicates the presence of interstellar neutral hydrogen that has charge exchanged with energetic stellar wind protons (Wood et al. 2005a). This shows that the star is embedded in an interstellar cloud containing neutral hydrogen. The absence of this blueshifted astrospheric absorption could be explained by the star being embedded in fully ionized interstellar gas or the sight line being near the downwind direction. Table 7 lists the stars within 10 pc for which astrospheric absorption has been detected or definitely not detected. In the 15 cases where there are detections of blueshifted astrospheric absorption, the star must be located within the cloud identified in the sight line. In three cases where there are more than one cloud, there is uncertainty as to which of the two or three clouds envelopes the star, but one of them must. For the eight cases where no astrospheric absorption has been reported, it is likely that no detected cloud extends as far as the star.

Table 7. Fill Factors for Sight Lines to Stars within 10 pc of the Sun

HDStar l b d (pc)Numberlog N(H i)FillingCloudsAstro-Ref.
 Name   Clouds Fractionin Frontsphere 
128620 α Cen ABC315.7−0.71.35117.610.49GYes1
GJ 699Barnard's star31.014.11.83117.720.46GNo3
95735GJ 411185.165.42.55117.840.44(LIC)No3
48915SiriusAB227.2−8.92.64217.4,17.20.25(LIC), (Blue)8
22049 epsilon Eri195.8−48.13.22117.880.32LICYes4
217987GJ 8875.1−66.03.29118.100.62LICYes2
20109161 Cyg A82.3−5.83.50217.8,17.80.58Eri, AqlYes8
61421Procyon213.713.03.51217.9,17.60.55LIC, Aur8
1326GJ 15A116.7−18.43.56118.020.48(LIC), (Hya)Yes2
209100 epsilon Ind336.2−48.03.64117.950.45(LIC)Yes5
10700 τ Cet173.1−73.43.65118.010.44LICNo1
GJ 273212.310.43.79217.86,17.780.57LIC, Aur2
GJ 191Kapteyn's star250.5−36.03.931?17.980.39Blue3
239960AGJ 860A104.7−0.04.01117.780.24(Eri)Yes2
GJ 388AD Leo216.554.64.97118.470.96(LIC), (Leo)No1
2696540 Eri A200.8−38.14.98117.80.21LICNo6
GJ 873.EV Lac100.6−13.15.05117.970.32(Hyades)Yes1
16534170 Oph A29.911.45.08317.8,17.1,17.50.34G, (Aql), (Mic)Yes8
187642 α Aql47.1−8.95.13317.9,17.9,17.50.60Aql, Eri, (Mic)8
36395GJ 205206.9–19.45.70217.60,17.240.16(LIC), unassignedYes2
GJ 754352.4−23.95.91218.16,17.620.51Unassigned, (Aql)9
GJ 588332.712.15.92118.120.36G2
15588636 Oph A358.36.95.96117.850.17GYes7
GJ 285YZ CMi215.913.55.99217.89,17.450.29LIC, (Aur)Yes2
191408GJ 783A5.2−30.96.02118.310.55(Mic)11
2079482 Eri250.7−56.76.041?18.330.57G3
190248 δ Pav329.8−32.46.10217.82,17.550.18Unassigned, (Vel)9
79210GJ 338A164.942.76.33117.790.16LICYes2
152751GJ 644B11.021.16.50118.400.63(Mic)2
131156 ξ Boo A23.161.46.73117.920.19GemYes1
H iP86287GJ 68642.224.38.16118.280.38LIC9
11561761 Vir311.944.18.57118.010.19(NGP)Yes?11
39587 χ1 Ori188.5−2.78.66117.930.12LICNo1
192310GJ 78515.6−29.48.81317.96,17.90,16.170.32(Vel), (Mic), (LIC)?10
23249 δ Eri198.1−46.09.04117.880.14LICYes1
20630 κ1 Cet178.2−43.19.14217.5,17.50.11LIC, (Hyades)No8
197481AU Mic12.7−36.89.72118.360.26(Mic)No1

References. (1) Wood et al. (2005b); (2) Wood et al. (2021); (3) Youngblood et al. (2022); (4) Dring et al. (1997); (5) Wood et al. (1996); (6) Wood & Linsky (1998); (7) Wood et al. (2000); (8) Redfield & Linsky (2008); (9) Zachary et al. (2018); (10) Edelman et al. (2019); (11) This paper.

Download table as:  ASCIITypeset image

The one-dimensional extent of a cloud along a given line of sight is more complicated. Since the Sun is embedded in the outer region of the LIC, other clouds detected in the same sight line must lie beyond the LIC. It is likely that the G Cloud also lies in front of other clouds as most of the sight line to the nearest star, α Cen, is in the G Cloud with no detected absorption at the LIC cloud's velocity. If one assumes that n(H i) in a cloud is the same as in the LIC (about 0.20 cm−3; Slavin & Frisch 2008; Swaczyna et al. 2020), then a cloud's path length along the sight line is given by the ratio N(H i)/0.20. While neutral hydrogen column densities N(H i) can be measured with modest precision, there are no accurate methods for measuring n(H i) in clouds other than the LIC.

High-resolution spectra of the Mg ii and Fe ii interstellar absorption lines observed toward nearby stars, primarily M dwarfs, provide a tool for identifying the filling fraction of a cloud along the sight line to the star. Table 7 lists all of the stars within 10 pc of the Sun with measured N(H i) and the number of clouds identified in each sight line. For each sight line, we identify the clouds located in front of the star by matching two criteria: (1) that the measured interstellar radial velocities lie within 2 km s−1 of the velocities predicted from the cloud's velocity vector, and (2) the star lies within the cloud's morphological outline (Redfield & Linsky 2008). Cloud names are placed in parentheses when the star is located at or just beyond the cloud edge.

The cloud-identification process is somewhat subjective as the cloud morphologies are not precisely known given the limited number of stars that were used to construct these morphologies. However, in all but three of the 52 velocity components in the 37 sight lines to stars within 10 pc, the number of measured interstellar-velocity components is the same as the number of previously known clouds identified by the two-step procedure just described. For the sight line to GJ 15A one absorption component was observed but two clouds meet both selection criteria. This could result from both clouds being in the sight line, or errors in the cloud outlines as the star is located at the edge of both the LIC and Hyades clouds. The agreement between the number of interstellar-velocity components and the number of identified clouds for 14 of the 15 new sight lines that were not available when the cloud outlines were constructed by Redfield & Linsky (2008) provides strong support for the cloud outlines obtained at that time. The identification of a cloud near its edge is a strong determination that the cloud is actually in the sight line, since there is no alternative cloud that meets both criteria of measured radial velocity and fitting within the cloud boundary. We conclude that nearly all velocity components identified so far correspond to previously identified warm clouds.

3.12. Are the Warm Clouds in the CLIC Closely Packed or Widely Separated?

The only cloud in the CLIC for which the neutral hydrogen density has been measured is the LIC, or more precisely the outer edge of the LIC where the heliosphere resides. Measurements of the inflowing neutral helium consistently predict n(H i) ≈ 0.20 cm−3. Slavin & Frisch (2008) developed theoretical models for the LIC just outside of the heliosphere consistent with this neutral hydrogen density and other empirical constraints. For other clouds in the CLIC, there are measurements of the neutral hydrogen column density, N(H i), but not the number density, n(H i). It is tempting to assume initially that n(H i) is roughly the same in the other CLIC clouds as in the LIC. If so, then it is simple to determine the extension of a cloud along a sight line from the measured N(H i) assuming n(H i) = 0.20 cm−3.

We can now test this assumption with data for 37 stars within 10 pc of the Sun. Table 7 lists the neutral hydrogen column densities for each cloud detected in the sight lines to all stars within 10 pc of the Sun that have measured N(H i). The eighth column lists the filling fractions of the clouds along each sight line determined from the cloud extensions (assuming that n(H i) = 0.20 cm−3) divided by the distance to the star. The ninth column lists the cloud(s) in the sight line to the star; clouds in parenthesis are viewed through their edges. For sight lines with multiple clouds, the filling fraction is the sum for all of the clouds. Figure 14 plots the filling fractions for the 37 sight lines. The least-squares fit to the data (the solid curve in the figure) shows a clear decrease in filling factor with distance to the star. The Pearson correlation coefficient r = −0.45 confirms the decrease. There are several interesting results shown in this figure:

  • 1.  
    The value of n(H i) entering the heliosphere and astrospheres plays a critical role in both ionization and heating because n(H i) sets the charge exchange rate with protons throughout the astrosphere and thus the relative fraction of suprathermal ions compared to the stellar wind ions. A higher number of suprathermal ions leads to more energy dissipation at the termination shock and beyond (see Swaczyna et al. 2020). It is therefore critical to obtain realistic estimates of n(H i).
  • 2.  
    With the exception of AD Leo (see Section 3.13), all of the sight lines have filling fractions smaller than 0.63. This maximum filling factor could be explained either by a sparcely filled CLIC with highly ionized gas located between the clouds observed along the sight line, or with clouds that entirely fill the sight lines but with smaller densities than the LIC value of n(H i) = 0.20 cm−3, or by a combination of the two possibilities. The large sample of 37 relatively short sight lines with 53 absorption components and the detection of 15 astrospheres provides an opportunity to distinguish between these possibilities.
  • 3.  
    Sight lines to 11 of the 13 stars within 4 pc of the Sun have filling factors in the range 0.38 to 0.62. Also, an additional five more distant stars have filling factors that lie within this range. This result could be explained by the clouds having typical densities of n(H i) ≈ 0.10 cm−3 that fill the sight lines. The alternative explanation of sight lines to a large number of stars all being about 50% filled seems contrived and less likely. The short sight line with the smallest fill factor (0.25) is toward Sirius AB (2.64 pc). This is not surprising because Sirius B is a hot white dwarf with a surrounding H ii region of fully ionized hydrogen.
  • 4.  
    There are four stars within 4 pc of the Sun (epsilon Eri, GJ 887, GJ 15A, and epsilon Ind, ) that have only LIC absorption in their sight lines and also astrospheric H i Lyα absorption that requires that these stars are located inside of a cloud containing neutral hydrogen. The simplest model for these sight lines is that the LIC extends from the heliosphere to the astrosphere of each star and perhaps beyond. In this case, the mean filling factor toward these four stars is 0.47 and the mean neutral hydrogen density in the LIC is 0.094 cm−3. The density n(H i) = 0.20 cm−3 in the immediate environment of the Sun, therefore, appears to be unrepresentative of the mean neutral hydrogen density in the LIC.
  • 5.  
    Other clouds in the CLIC may have similar densities to n(H i) ≈ 0.10 cm−3 for the LIC. For example, the sight line to the nearest stellar binary system α Cen AB shows no evidence for LIC absorption, only G Cloud absorption and astrospheric absorption likely produced by neutral hydrogen in the G Cloud. After a minimal amount of LIC gas, the complete sight line to α Cen could be filled with G Cloud gas with n(H i) = 0.10 cm−3. We conclude that within 4 pc of the Sun the clouds appear to be tightly packed, and that typical densities are n(H i) ≈ 0.10 cm−3. However, a "Swiss cheese" model with high-density gas separated by voids cannot be ruled out by the available data.
  • 6.  
    Beyond 4 pc, there is a wider range of filling fractions with a clear trend of decreasing filling fraction with increasing distance, as shown by the least-squares linear fit shown in Figure 14. For the sight lines between 4 pc and 7 pc the filling fractions have a wide range (0.16–0.63), but the seven sight lines in the range 8–10 pc all have filling fractions less than 0.38, and four of the sight lines have filling fractions of about 0.15. Gry & Jenkins (2017) found that the warm gas filling factor in the sight line to α Leo (23.8 pc) is about 0.13. This trend of decreasing filling fraction with distance suggests that the CLIC clouds are becoming more widely separated with distance from the Sun, as was suggested by Redfield & Linsky (2008), and that the intercloud gas with fully ionized hydrogen is occupying a larger fraction of sight lines to the more distant stars. The intercloud gas may be the same as the gas that fills most of the Local Cavity (Linsky et al. 2019).

Figure 14.

Figure 14. Plot of the filling fraction of neutral hydrogen clouds along sight lines to stars vs. stellar distance. On the right side is the scale of neutral hydrogen density if the sight lines are completely filled with clouds. The left colored box identifies sight lines that would be filled by clouds with a mean density n(H i) ≈ 0.10 cm−3. The right colored box identifies sight lines with distant clouds that are widely separated. The four stars within 4 pc of the Sun with only LIC and astrospheric absorption are identified in the figure and are discussed in the text.

Standard image High-resolution image

Although the model just described fits the data very well, it has been suggested to explore an alternative model in which warm clouds fill the entire space well beyond 4 pc. If filling factors are close to unity along these longer sight lines, then one might expect that the thermal plus turbulent pressure to be roughly constant despite the inhomogeneous temperatures and turbulent velocities. Values of variable magnetic and other pressure terms are not available for individual sight lines. We therefore looked for correlations of the mean neutral hydrogen density 〈n(H I)〉 = N(H I)/d with thermal pressure, proportional to T, and with the sum of thermal and turbulent pressure, proportional to ${kT}+1.4m({\rm{H}}){v}_{\mathrm{turb}}^{2}$. We selected 16 sight lines, listed in Table 7, that have only a single cloud. In Figure 15 there is a Pearson correlation coefficient, r = 0.24, for the correlation of 〈n(H I)〉 with thermal pressure, and a correlation coefficient, r = 0.26, for the correlation of 〈n(H I)〉 with the sum of thermal and turbulent pressures. The slopes of the least-squares fits are less than their errors in both plots. The larger correlation coefficient for filling factor with distance in Figure 14, r = −0.45, supports the previous model of warm clouds completely filling space out to 4 pc with decreasing filling fraction beyond.

Figure 15.

Figure 15. Plots of the mean density assuming filling factor of unity vs. temperature (left) and the sum of thermal and turbulent pressure (right). The solid lines are least-squares linear fits to the data.

Standard image High-resolution image

3.13. Is There a Shock in the Sight Line to AD Leo?

The short (4.97 pc) sight line to AD Leo is interesting for several reasons. The Galactic coordinates of AD Leo (l = 216fdg5, b = +54fdg6) place it at the edge of the LIC, Leo, and north Galactic pole (NGP) clouds (Redfield & Linsky 2008), but its interstellar radial velocity of 13.13 km s−1 is inconsistent with the predicted radial velocities of the LIC (7.23 km s−1), Leo (9.21 km s−1) and NGP (16.22 km s−1) clouds, using the LISM Dynamical Model Kinematic Calculator. 4 As a result, AD Leo is not assigned to any known cloud and the high H i column density may result from the sight line passing through a shock. Since the LIC and Leo clouds have essentially the same velocity amplitudes, the shock would not be from these two clouds. Instead, the NGP cloud has a velocity vector amplitude 13 km s−1 different from the LIC and Leo clouds and thus could be responsible for the shock.

Gry & Jenkins (2017) found interstellar absorption toward the star Regulus (α Leo) at 8.8 ± 0.2 km s−1, which they assigned to the LIC, and a second component at 14.4 ±0.1 km s−1. Since Regulus and AD Leo are separated by only about 10° and the radial velocity of the second component is similar to that seen toward AD Leo, it is likely that these two components are formed in the same unassigned structure in front of AD Leo and thus closer than 4.97 pc. Although the Local Leo Cold Cloud (LLCC; Peek et al. 2011) is in the same direction as Regulus (and AD Leo) with a similar radial velocity (9.3 km s−1), Gry & Jenkins (2017) argue that the LLCC lies beyond Regulus at a distance of 33.5 ± 11.3 pc. Thus the anomalously large value of N(H i) in front of AD Leo is not the LLCC, but some other perhaps very interesting feature.

If the high value of N(H i) in front of AD Leo is produced in a shock between the NGP and either the LIC or Leo clouds, then the radial velocity of the shock should be intermediate between the NGP and the other two clouds. The halfway radial velocity between the NGP and LIC is 11.7 km s−1 and between the NGP and Leo clouds is 12.7 km s−1. Both velocities are consistent with the observed interstellar velocity of 13.13 km s−1 toward AD Leo. This spatial agreement between AD Leo and the cloud interfaces and the velocity agreement provides evidence for a shock in the sight line to AD Leo. This would be a second shock in the LISM, the first being the Cetus ripple discovered by Gry & Jenkins (2014) located mostly in the southern Galactic hemisphere.

4. Discussion

4.1. The Development of Empirical Local Interstellar Medium Studies

Empirical studies of the local ISM can be viewed as having proceeded through three stages driven by the increasing availability of high-resolution UV spectra and the constraint of matching the properties of the plasma flowing from the LISM into the heliosphere. The first stage consisted of studies of individual sight lines to nearby bright stars, first by analyzing ground-based observations of the Ca ii H and K lines, and then by analyzing UV spectra from the Copernicus satellite and the HRS instrument on HST. These studies provided measurements of radial velocities, temperatures, turbulent broadening, hydrogen and metal column densities, and electron densities of interstellar matter in these sight lines. Subsequently, Redfield & Linsky (2004b) measured the properties for 50 velocity components along the sight lines to 29 stars using high-resolution spectra from the STIS instrument on HST. These studies could not determine where along the sight line the absorption occurs or whether these properties are homogeneous or the mean of variations along the sight line.

The recognition that the flow of interstellar gas is in the form of comoving structures, now called clouds, is the second stage in the empirical study of the LISM. It began with the discovery by Crutcher (1982) that radial velocities in many sight lines in directions away from the Galactic Center are consistent with a coherent flow from the direction of the Scorpio-Centaurus Association, and that the properties of the gas in this flow are consistent with the properties of neutral helium atoms flowing into the heliosphere from the LISM (Witte et al. 1993). The region from which this flow originates is now called the LIC. Subsequently, Lallement & Bertin (1992) found that the flow of gas observed for sight lines in the Galactic Center direction are consistent with a different flow vector that they named the Galactic cloud, and which is now called the G Cloud. Lallement et al. (1994) noted that the sight line to Sirius A shows a second velocity feature at −5.7 ± 0.2 km s−1 relative to the LIC absorption. A similar blueshifted absorption component in the direction of epsilon CMa observed by Gry & Jenkins (2001) confirmed that this extra absorption is from a third cloud called the Blue Cloud.

With the accumulated database of 270 radial velocity measurements toward 157 stars within 100 pc of the Sun, Redfield & Linsky (2008) identified 15 velocity vectors. The validity of these clouds and their morphologies was confirmed by Malamut et al. (2014), who found that nearly all of the newly observed velocity components that lie within the morphologies of clouds previously identified by Redfield & Linsky (2008) have radial velocities consistent with these cloud vector velocities. Further support for the multicomponent model comes from the scintillation of radio emission from point-source quasars (Linsky et al. 2008). In the present paper we find that the radial velocities of 49 out of 52 velocity components toward stars within 10 pc have radial velocities consistent with the clouds in their directions. Although the multicloud model fits essentially all of the available data, the assumption of discrete clouds with constant internal flows and finite edges may be unrealistic. An alternative model proposed by Gry & Jenkins (2014) in which the LISM consists of one cloud with internal velocity gradients filling all of space within 9 pc may be more realistic, although it does not fit the data as well as the multicomponent model (Redfield & Linsky 2015).

The third stage in the development of empirical LISM studies, beginning with this paper, is the recognition that the temperatures and turbulent velocities within the LIC and other clouds are not homogeneous but have a wide range of values. Whether or not this distribution is random remains to be seen. A future stage in understanding of the LISM would be the identification of the physical causes responsible for these variable parameters on the basis of observations. A related question is whether the clouds completely fill space within about 4 pc, as discussed in Section 3.12, or whether an intercloud medium separates the clouds especially at larger distances. Breitschwerdt et al. (2000) proposed that the LIC and other warm clouds are produced by fragmentation due to hydromagnetic Rayleigh–Taylor instabilities that occur where the Local Bubble and Loop I interact. This formation mechanism would be consistent with the clouds being isolated structures separated from other clouds by an ionized intercloud medium.

4.2. The Development of Theoretical Interstellar Medium Studies

In parallel with the empirical studies, theoretical models and physically based simulations have developed in two stages. The first stage involved models that included many heating and cooling processes, but assumed that there is energy and pressure balance in an assumed quiescent and static ISM. Theoretical models of a multicomponent ISM (Field et al. 1969; Wolfire et al. 1995, 2003) assume that the balance between heating and cooling processes plays the critical role in identifying the likely temperature-pressure structures in the ISM. In these models, warm interstellar gas is usually modeled as two stable phases: the warm neutral medium (WNM), consisting of neutral hydrogen and other species, and the warm ionized medium (WIM) in which hydrogen is fully ionized. The WNM and WIM can coexist at the same thermal pressure given the model assumptions. The WIM is often called an H ii region ora Strömgren sphere surrounding a star emitting strong EUV radiation (see Linsky et al. 2019).

Since there is strong evidence that near the heliosphere the LIC is partially ionized, ne /nH = 0.07 cm−3/0.195 cm−3 = 0.35 (Slavin & Frisch 2008), neither of the two warm models may provide a useful prototype for the properties of the LIC and other nearby clouds. Our data on temperatures in the LIC and nearby partially ionized clouds can test these models.

In the second stage, numerical simulations of the ISM powered by supernova explosions and winds and radiation from hot stars predict a very different ISM. The simulations of de Avillez & Breitschwerdt (2005) and de Avillez & Breitschwerdt (2012), for example, show that dynamical phenomena resulting from shock waves and instabilities create a time-dependent nonequilibrium ISM in which there are no steady-state phases and all parameters (temperatures, turbulent velocities, densities, and magnetic fields) have a wide range of values both spatially and over time. Also, highly nonlinear heating and cooling processes mean that time-independent energy balance is not realistic. In these simulations more than half of the mass is in thermally unstable phases predicted by the steady-state models. Are these simulations relevant for describing the LISM embedded inside an old supernova remnant, or are the steady-state theoretical models of a more quiescent LISM a better approximation? The diverse properties of the LIC and the nearby warm clouds provides an opportunity to begin testing these very different models.

4.3. Tests of Theoretical Models

One of the predictions of the steady-state theoretical models is that temperatures in both the WNM and WIM lie in the thermally stable range 5040–8310 K (Wolfire et al. 2003) with wider limits depending on the rates of heating and cooling. Lower temperatures should be unstable due to rapid cooling to very low temperatures becoming CNM gas. By contrast, the probability distribution function for temperatures is roughly flat between 100 K and 106 K in the Breitschwerdt & de Avillez (2021) simulations. We measured temperatures in a number of clouds with T > 10,000 K, the highest temperature with small uncertainty being $T=12,{050}_{-790}^{+820}$ K for the sight line to V368 Cep through the LIC. These high temperatures are probably inconsistent with steady-state energy balance in the WNM and WIM models and the rapid increase of Lyα cooling with increasing temperature.

Table 8 lists the sight lines through clouds with low interstellar gas temperatures. There are 13 velocity components in 10 sight lines with temperatures below 3500 K, but many have large uncertainties. Four velocity components have temperatures including 1σ positive errors that lie below 3500 K: HD 72905 (π1 UMa), HD 129333 (EK Dra), and HD 220657 (υ Peg). These stars do not lie in similar directions, but both EK Dra and υ Peg are at the edges of clouds and the sight line to π1 UMa passes close to the center of the LIC. These velocity components provide evidence for interstellar matter somewhat cooler than T < 3500 K, well below the nominal temperature regime for WNM models. The low temperatures are distributed among seven known clouds and two sight lines without known clouds. In most cases the sight lines pass through the edges or just outside of clouds. We note that the sight line to 70 Oph includes low temperatures through three clouds, and the sight line to υ Peg contains low temperatures through two clouds. These two stars are well separated in Galactic coordinates. However, the sight lines to SAO 28753, η UMa, and EK Dra are in similar directions, suggesting that these low temperatures could have a common origin. Since two of these identifications are unassigned to any known cloud and the sight line to the third (EK Dra) passes through the edge of or just outside of the LIC, the low temperatures may occur in interstellar matter outside of partially ionized clouds, perhaps in the WNM. The LLCC, located between 11.3 pc and 24.3 pc from the Sun according to Peek et al. (2011), has a temperature of 15–30 K, which is definitely CNM. None of the low-temperature sight lines listed in Table 7 is near the LLCC (centered at l = 222°, b = 44°).

Table 8. Low-temperature Sight Lines

HDName l b d (pc)v T ξ Ref.Cloud
48915 α CMa227.2−8.92.612.70 ${3000}_{-1000}^{+2000}$ 2.7 ± 0.32Blue
GJ 873EV Lac100.6−13.15.04.44 ± 0.43 ${3030}_{-610}^{+670}$ ${2.21}_{-0.23}^{+0.22}$ 1(Hyades)
16534170 Oph29.911.45.1−26.50 ± 0.07 ${2700}_{-2300}^{+3000}$ ${3.64}_{-0.44}^{+0.42}$ 2G
16534170 Oph29.911.45.1−32.53 ± 1.30 ${1700}_{-1700}^{+2100}$ 3.3 ± 1.12(Oph)
16534170 Oph29.911.45.1−43.34 ± 0.923300 ± 21002.31 ± 0.372(Aql)
72905 π1 UMa150.635.714.512.91 ± 0.43 ${2450}_{-660}^{+740}$ ${2.47}_{-0.11}^{+0.10}$ 1LIC
43162GJ 3389230.9−18.516.717.1 ± 0.43 ${3200}_{-2330}^{+2680}$ ${5.76}_{-0.26}^{+0.25}$ 1LIC
116956SAO 28753113.759.521.77.7 ${2540}_{-1830}^{+2450}$ ${1.66}_{-1.66}^{+1.21}$ 1Unassigned
120315 η UMa100.765.330.92.6 ± 3.4 ${0}_{-0}^{+4400}$ ${5.6}_{-1.1}^{+0.9}$ 2Unassigned
129333EK Dra105.549.034.6−2.43 ± 0.32 ${2580}_{-250}^{+280}$ ${3.13}_{-0.05}^{+0.05}$ 1(LIC)
93497 μ Vel283.008.635.9−4.38 ± 0.43 ${3470}_{-590}^{+580}$ ${3.12}_{-0.10}^{+0.11}$ 1G
220657 υ Peg98.6−35.453.18.8 ± 1.2 ${1000}_{-1000}^{+1900}$ ${3.46}_{-0.63}^{+0.61}$ 2(Hyades)
220657 υ Peg98.6−35.453.11.73 ± 0.39 ${1700}_{-900}^{+1100}$ 3.93 ± 0.222(Eri)

References. (1) This paper; (2) Redfield & Linsky (2004b).

Download table as:  ASCIITypeset image

4.4. Possible Causes for the Diversity of Cloud Temperatures

The observed diversity of temperatures inside the LIC and in other partially ionized clouds raises the question of what physical processes could be responsible. Localized heating by hydromagnetic shocks is an obvious candidate that needs to be pursued with high-spatial-resolution observations. Another possibility is a change in the energy balance on small spatial scales resulting from a change in the local heating or cooling rates driven by local changes in the ionization. Cooling per unit volume by the emission of optical and UV photons following electron collisional excitation of atoms and ions is proportional to the product ne nH i (Draine 2011). If the dominant heating process is from photoelectrons emitted by dust grains following absorption of UV photons, then the heating rate per atom is proportional to the dust density and thus the heating rate per unit volume is proportional to nH i 2. If these two processes dominate the energy balance, then ne plays a critical role. Higher electron densities lead to enhanced cooling and thus lower temperatures, while lower electron densities lead to decreased cooling and thus higher temperatures. Support for this prediction comes from the Gry & Jenkins (2001) analysis of the epsilon CMa sight line, where ne is lower in the Blue Cloud than the LIC but the temperature of the Blue Cloud is higher.

All previous analyses of absorption lines in the LISM and the present work have assumed that the component of line broadening that depends on mass is thermal, so that line widths measure gas temperatures. Studies of the heliosphere indicate that small-scale nonthermal processes such as shocks and charge exchange produce pickup ions that create a plasma with both thermal and suprathermal velocities. Suprathermal velocities are most easily detected in lines of low-mass atoms such as D i (m = 2), because turbulent broadening dominates over thermal broadening for high-mass ions such as Mg ii (m = 24). Thus what we have called temperature may in fact be a combination of thermal and suprathermal velocities, and the observed inhomogeneous "temperatures" may be due in part to spatial differences in the suprathermal broadening. Also, the different timescales for ionization and recombination for warm gas in the ISM cause the plasma to be out of equilibrium such that the electron temperature responsible for line broadening can be very different from a "temperature" characterizing ionization balance (e.g., de Avillez & Breitschwerdt 2012).

Small-scale variations in the thermal and suprathermal velocity structure can be maintained for a long time by inhomogeneous magnetic fields, since the Larmor radius of a 7000 K thermal electron in a 3 μG magnetic field is 0.2 km and for a thermal proton is 350 km. Thus inhomogeneous structures smaller than 5000 au can be maintained by random magnetic fields much smaller than 3 μG even for protons with energies much larger than thermal.

5. Conclusions

With the increase in the number of sight lines with analyzed interstellar properties, we have explored the inhomogeneity of these parameters within the LIC and other clouds in the CLIC with the following results:

  • 1.  
    Temperatures and turbulent velocities within the LIC and other nearby warm clouds have a wide range of values. The distributions appear to be random and can be fit with Gaussians. Within the LIC, temperatures range between 2450 K and 12,900 K and turbulent velocities range between 0.0 and 4.4 km s−1. Since we measure average quantities along a sight line, the temperature and turbulent velocity measurements could underrepresent the true inhomogeneities. The presence of temperature inhomogeneities implies density inhomogeneties as well. Previous methods of characterizing the properties of the LIC and other clouds by mean temperatures and turbulent velocities are not realistic and likely conceal important physical processes producing the inhomogeneities.
  • 2.  
    Comparison of temperature differences between pairs of sight lines across the LIC indicate that the angular scale for inhomogeneous temperatures is less than 2fd2. This angular scale corresponds to a distance scale of ≤5100 pc, a distance that the Sun will traverse in less than 1000 yr. The size and shape of the heliosphere will change when the Sun encounters changes in the ISM.
  • 3.  
    Temperatures and turbulent velocities do not show any statistically significant trends with respect to the distance to the background star, Galactic coordinates, hydrogen column density, or angle relative to the direction of the magnetic field or interstellar gas flowing into the heliosphere. The absence of a trend with stellar distance allows one to compare the properties of clouds in sight lines to stars located over a wide range of distances. The absence of trends with angle from the center of the LIC means that the properties of the center and the edge of the LIC are similar. The absence of trends in the temperature and turbulent velocity with angle relative to the strong EUV radiation source epsilon CMa indicates that photoionization by multiple sources of EUV and UV radiation must be considered in determining the heating rates within the LIC and other clouds.
  • 4.  
    For 34 of the 37 sight lines to stars within 10 pc of the Sun, there is a unique match of each velocity component and direction with a previously identified warm cloud in the CLIC. If the density of neutral hydrogen is the same in all CLIC clouds as in the LIC near the Sun, nH i = 0.20 cm−3, then none of these sight lines is completely filled with the identified clouds, and the clouds are separated by a fully ionized intercloud medium. If our conclusion is that nH i ≈ 0.10 cm−3 is typical for nearby clouds, then all of the sight lines within 4 pc are completely filled with warm clouds without separation by an ionized intercloud medium, but at larger distances the clouds are more widely separated. The question of the distance that the CLIC clouds extend into space from the Sun may be partially answered by this model.
  • 5.  
    A critical test of the neutral hydrogen density of the LIC is provided by the four stars within 4 pc of the Sun with sight lines that have only LIC absorption and astrospheres that have Lyα absorption requiring that partially ionized hydrogen surrounds these stars. The most sensible model for these sight lines is that LIC gas completely fills these sight lines. The mean fill factor for these sight lines is 0.47, indicating that the mean n(H i) = 0.094 cm−3. We conclude that the mean neutral hydrogen number density in the LIC and perhaps other warm clouds in the CLIC is about 0.10 cm−3, and that the higher density of 0.20 cm−3 surrounding the heliosphere is anomalous, perhaps because of overlap with G Cloud gas as proposed by Swaczyna et al. (2022).
  • 6.  
    Since warm gas in the nearby clouds is about one-third ionized, the clouds may not be described by steady-state theoretical models of the WNM and WIM, which assume either completely neutral or highly ionized gas with different heating and cooling processes. Both models predict maximum temperatures of about 10,000 K controlled by the rapidly rising cooling of Lyα emission with increasing temperature. We find a number of sight lines with T > 10,000 K, but none with temperatures in excess of about 13,000 K. We find 10 sight lines with relatively low temperatures, T ≤ 3500 K, which is inconsistent with the predictions of steady-state theoretical models that the temperature regime between 5000 K and 300 K is unstable at constant pressure. These cool temperature sight lines traverse seven known clouds and two directions where no clouds are presently known. Three of the sight lines are in similar directions, suggesting that the low temperatures have a common origin. Since two of these sight lines do not traverse known clouds and the direction of the third may be outside of the LIC, we suggest that these regions of cool gas could be examples of CNM.
  • 7.  
    We list several possible processes that could be responsible for the local variations in temperature within the LIC and other partially ionized clouds. Heating by shocks is an obvious candidate, but we consider two other possibilities. One possibility is a model in which the dominant heating process in the partially ionized clouds is from energetic photoelectrons emitted by dust grains following the absorption of stellar UV photons, and the dominant cooling process is by electron excitation of atoms and ions followed by emission of optical and UV photons. The heating rate is proportional to nH i and the cooling rate is proportional to the product ne nH i . In this case, the electron density and thus the relative ionization plays a critical role as increases in ne produce more cooling, thereby lowering the temperature. Another possibility is that what is measured as "temperature" may be a measure of both thermal and suprathermal velocities with the nonthermal component produced by charge exchange and other processes. The diversity in the measured "temperatures" could be produced in part or entirely by spatial variations in the suprathermal component.

We acknowledge support from the NASA Outer Heliosphere Guest Investigators Program to Wesleyan University and the University of Colorado for grant No. 80NSSC20K0785. The spectra used in this paper were obtained by the HST, primarily from the SNAP programs 10236, 11568, 13332, 13658, 14084, and 16487 but also from programs 13650 and 15071. The data were extracted from MAST. We thank Brian E. Wood for his comments on the AD Leo data. We especially wish to thank the referee for many excellent suggestions and a very careful reading of the manuscript.

Facility: HST(STIS), HST(COS), Voyager I, Voyager II.

Appendix: Parameters for All Sight Lines (Past and Present) to 100 pc

Table A1 is a compilation of data for the 130 sight lines to stars within 100 pc for which there are published high-resolution UV spectra relevant to the LISM.

Table A1. Parameters for All Available Sight Lines within 100 pc

HD or GJStar (Cloud) d (pc)Gal. long.Gal. lat.Temperature+/- errorTurb. Velocity+/- errorlog N(H i)
128620Alpha Cen A (G)1.3315.7−0.75100.01150.01.210.4117.61
128621Alpha Cen B (G)1.3315.7−0.75500.0325.01.370.3717.61
GJ 699Barnard's star (G)1.8331.014.117.72
95735GJ 411 (LIC)2.55185.165.417.84
48915Sirius A (LIC)2.6227.2−8.98000.0750.00.50.317.2
48915Sirius A (Blue)2.6227.2−8.93000.01500.02.70.317.2
48915Sirius B (LIC)2.6227.2−8.98000.0750.00.50.317.6
48915Sirius B (Blue)2.6227.2−8.93000.01500.02.70.317.6
22049Epsilon Eri (LIC)3.2195.8−48.17410.0845.02.030.4317.88
21798GJ 887 (LIC)3.295.1−66.06000.01450.01.650.9118.10
61421Procyon (LIC)3.5213.713.06710.0645.01.210.4017.9
61421Procyon (Aur)3.5213.713.06710.0645.01.210.4017.6
20109161 Cyg A (Aql)3.582.3−5.86850.0880.02.080.6417.8
20109161 Cyg A (Eri)3.582.3−5.86850.0880.02.080.6417.8
1326GJ 15A (Hya)3.56116.7−18.47800.01150.02.810.2118.12
209100Epsilon Ind (LIC)3.64336.2−48.08340.0445.01.970.0918.10
10700Tau Cet (LIC)3.65173.1−73.4<670018.01
GJ 273GJ 273 (LIC)3.79212.310.48620.0350.00.01.3217.86
GJ 273GJ 273 (Aur)3.79212.310.48620.0345.00.01.5917.78
33793Kapteyn's star (Blue)3.93250.5−36.017.98
239960GJ 860A (Eri)4.01104.7−0.04780.01930.02.240.4617.78
GJ 388AD Leo (LIC)4.97216.554.6<1230018.47
2696540 Eri A (LIC)4.98200.8−38.18120.0450.00.50.8517.8
GJ 873EV Lac (LIC)5.05100.6−13.13030.0640.02.210.2217.97
16534170 Oph (G)5.0829.911.42700.02650.03.640.4117.8
16534170 Oph (Aql)5.0829.911.43300.02100.02.310.3717.1
16534170 Oph (Mic)5.0829.911.41700.01900.03.31.117.5
187642Alpha Aql (Aql)5.1347.1−8.912,600.02400.00.630.7617.9
187642Alpha Aql (Eri)5.1347.1−8.912,300.02100.00.001.2017.9
187642Alpha Aql (Mic)5.1347.1−8.912,500.02550.01.41.0517.5
36395GJ 205 (LIC)5.70206.9−19.44500.01200.03.640.2717.60
36395GJ 205 (Unassigned)5.70206.9−19.44200.0680.02.370.2417.24
GJ 754GJ 754 (Unassigned)5.91352.4−23.918.16
GJ 754GJ 754 (Aql)5.91352.4−23.917.62
GJ 588GJ 5885.92332.712.16730.0710.03.370.3218.12
15588636 Oph A (G)5.96358.36.95870.0560.02.330.4817.85
GJ 285YZ CMi (LIC)5.99215.913.58930.0260.00.01.1817.89
GJ 285YZ CMi (Aur)5.99215.913.58790.0230.01.590.6717.45
191408GJ 783A (Mic)6.015.2−30.97700.01500.03.050.8618.31
2079482 Eri (G)6.04250.7−56.718.33
190248Delta Pav (Unassigned)6.11329.8−32.48680.0760.00.02.1817.82
190248Delta Pav (Vel)6.11329.8−32.49310.09100.02.441.7217.55
79210GJ 338A (LIC)6.33164.942.76650.0995.02.290.5818.02
152751GJ 644B (Mic)6.5011.021.118.40
131156Xi Boo A (Gem)6.7323.161.45310.0830.01.680.2317.92
GJ 686GJ 686 (LIC)8.1642.224.318.28
11561761 Vir (NGP)8.53311.944.16600.04600.01.60,1.2018.01
39587Chi Ori (LIC)8.7188.5−2.77000.0705.02.380.1617.93
192310GJ 785 (Vel)8.8015.6−29.48600.01900.03.31.2517.96
192310GJ 785 (Mic)8.8015.6−29.49900.02150.02.21.5517.90
192310GJ 785 (LIC)8.8015.6−29.46900.02450.01.31.4516.17
23249Delta Eri (LIC)9.04198.1−46.03650.01160.02.250.2117.88
20630Kappa1 Cet (LIC)9.14178.2−43.15200.01800.02.640.3017.5
20630Kappa1 Cet (Hya)9.14178.2−43.13600.02550.02.170.4317.5
20630Kappa1 Cet (Unassigned)9.2178.2−43.15800.02700.01.480.9217.4
197481AU Mic (Mic)9.7212.7−36.88700.01200.04.300.9318.36
GJ 678.1AGJ 678.1A (Oph)10.1228.620.510,72043000.001.9217.00
GJ 678.1AGJ 678.1A (LIC)10.1228.620.585403,0200.001.3418.57
62509Beta Gem (LIC)10.3192.223.49000.01550.01.670.3018.0
62509Beta Gem (Gem)10.3192.223.46100.02850.01.930.6917.8
17925EP Eri (LIC)10.4192.1−58.38100.01300.02.461.4518.0
33262Zeta Dor (Blue)11.7266.0−36.77700.02200.02.340.4317.8
33262Zeta Dor (Dor)11.7266.0−36.77000.03250.05.470.4018.1
37394GJ 211 (LIC)12.3158.411.95600.02350.01.000.89518.21
34029Capella (LIC)12.9162.64.66700.01350.01.680.3618.2
9826Ups And (Unassigned)13.5132.0−20.710400.01950.00.02.417.78
9826Ups And (Hyades)13.5132.0−20.74000.02500.01.81.117.55
9826Ups And (LIC)13.5132.0−20.76500.02800.01.31.017.34
166HR 8 (LIC)13.7111.3−32.84520.02000.04.390.3418.46
72905Pi1 UMa (LIC)14.5150.635.72450.0700.02.470.1118.32
142373Chi Her (Mic)15.967.750.34700.01150.01.360.3618.21
432Beta Cas (LIC)16.7117.5−3.39760.0840.00.01.118.2
43162GJ 3389 (LIC)16.7230.9−18.53200.02500.05.760.2617.97
165185GJ 702.1 (LIC)17.2356.0−7.38610.03750.00.001.7618.09
82443DX Leo (LIC)17.7201.246.18540.0850.01.781.7717.7
206860NN Peg (Unassigned)17.969.9−28.37100.02600.01.41.017.35
206860NN Peg (Eri)17.969.9−28.39600.02400.02.110.6117.79
206860NN Peg (LIC)17.969.9−28.36800.02600.00.800.8918.14
82558LQ Hya (G)18.3244.628.45700.06000.02.521.2618.8
22014V368 Cep (LIC)19.7118.516.912,050.0805.00.01.018.0
11443Alpha Tri (LIC)19.7138.6−31.47700.02850.00.01.718.1
11443Alpha Tri (Hyades)19.7138.6−31.48900.03650.01.31.517.8
29139Alpha Tau (Aur)20.4181.0−20.217.8
203554SAO 254993 (Vel)20.8324.9−38.918.8
116956SAO 28753 (LIC)21.7113.759.55160.03050.01.791.0318.10
116956SAO 28753 (Unassigned)21.7113.759.52540.02140.01.661.4517.68
59967GJ 3446 (Blue)21.8250.5−9.012,400.02800.00.001.918.41
1405PW And (LIC)21.9114.6−31.411,300.01850.00.01.618.1
97334GJ 417 (LIC)22.7184.367.311,000.02700.00.01.418.11
106516HR 4657 (Leo)22.4288.551.5<27,50018.6
128987SAO 158720 (Unassigned)23.8337.539.2<12,40018.1
87901Alpha Leo (LIC)23.8226.448.96000.0600.01.780.1018.18
87901Alpha Leo (NGP)23.8226.448.95990.0700.01.850.1917.67
73350SAO 136111 (LIC)24.3232.120.07900.03200.00.681.6717.8
222107Lambda And (LIC)25.8109.9−14.5<110018.5
22468HR 1099 (LIC)29.0184.9−41.67900.01500.01.180.4717.9
22468HR 1099 (Hyades)29.0184.9−41.68800.0950.00.00.917.6
22468HR 1099 (Unassigned)29.0184.6−41.67100.01400.02.300.2517.2
4128Beta Cet (Mic)29.4111.3−80.712,400.02800.02.290.4416.9
4128Beta Cet (Cet)29.4111.3−80.76300.02900.01.310.7618.5
120315Eta UMa (NGP)30.9107.765.38900.02400.01.340.2817.9
120315Eta UMa (Unassigned)30.9107.765.30.04400.05.61.016.6
137759Iota Dra (NGP)31.094.048.617.88
HZ 43 (NGP)32.054.184.27500.02050.01.71.2518.0
82210DK UMa (LIC)32.4142.538.94220.0446.01.860.1718.10
129333EK Dra (LIC)34.6105.549.02580.0265.03.130.0518.24
93497Mu Vel (G)35.9283.08.63470.0585.03.120.1118.67
62044Sigma Gem (Gem)37.5191.223.37200.01100.00.01.117.7
62044Sigma Gem (Unassgned)37.5191.223.38600.01600.02.460.4517.9
28568SAO93981 (LIC)45.2180.5−21.418.0
128987KU Lib (Gem)42.3331.030.26580.04090.03.00.1117.74
197890BO Mic (Vel)44.46.2−38.310,600.02700.03.481.8718.3
28205V993 Tau (LIC)47.8180.4−22.418.0
28033SAO 76609 (LIC)48.4175.4−18.9<12,90018.2
V471 Tau (LIC)47.7172.5−27.9<11,00018.2
209458V376 Peg (Eri)48.176.8−28.58500.03065.00.03.0218.4
220657Upsilon Peg (Eri)53.198.6−35.41700.01000.03.930.2217.9
220657Upsilon Peg (Hyades)53.198.6−35.41000.01450.03.40.6217.4
220657Upsilon Peg (Unassigned)53.198.6−35.43600.03700.01.71.117.4
32008HR 1608 (LIC)52.9209.6−29.417.8
32008HR 1608 (Unassigned)52.9209.6−29.4<11,50017.7
203387Iota Cap (LIC)66.133.6−40.812,900.03550.01.580.7317.9
203387Iota Cap (Mic)66.133.6−40.811,700.03850.03.820.4118.1
203387Iota Cap (Unassigned)66.133.6−40.85500.08050.03.70.918.4
G191-B2B (LIC)68.8155.97.16200.01350.01.780.4618.2
G191-B2B (Hyades)68.8155.97.14400.02600.03.270.3817.4
23189GD 153 (NGP)70.5317.384.77000.02850.01.21.1517.9
Feige 24 (LIC)74.4166.0−50.218.1
Feige 24 (Unassigned)74.4166.0−50.218.1
11181231 Com (NGP)94.2115.089.68200.01200.00.01.018.0

Note. Column 1: The HD catalog numbers for the star at the end of the sight line when available. When the star is too faint to be included in the HD catalog, the GJ star number is listed instead. The column is blank for the few stars without HD or GJ catalog numbers. Column 2: The star name with the interstellar cloud name in parenthesis. If no known cloud has been identified for the absorber, the term "unassigned" is used. Column 3: Distance to the star in parsecs. Columns 4 and 5: Galactic longitude and latitude. Columns 6 and 7: Measured temperature and error for the cloud in this sight line. No available data is indicated by "...". Columns 8 and 9: Measured turbulent velocity and error for the cloud in this sight line. No available data is indicated by "...". Column 10: Log of the neutral hydrogen column density through the cloud.

A machine-readable version of the table is available.

Download table as:  DataTypeset images: 1 2 3

Footnotes

  • All of the data presented in this paper were obtained from the Mikulski Archive for Space Telescopes (MAST) at the Space Telescope Science Institute. The specific observations analyzed can be accessed via https://doi.org/10.17909/gabn-m136. STScI is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. Support to MAST for these data is provided by the NASA office of Space Science via grant No. NAG5-7584 and by other grants and contracts.

  • 4  
Please wait… references are loading.
10.3847/1538-3881/ac816b