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

ICE CHEMISTRY IN STARLESS MOLECULAR CORES

Published 2015 June 18 © 2015. The American Astronomical Society. All rights reserved.
, , Citation J. Kalvāns 2015 ApJ 806 196 DOI 10.1088/0004-637X/806/2/196

0004-637X/806/2/196

ABSTRACT

Starless molecular cores are natural laboratories for interstellar molecular chemistry research. The chemistry of ices in such objects was investigated with a three-phase (gas, surface, and mantle) model. We considered the center part of five starless cores, with their physical conditions derived from observations. The ice chemistry of oxygen, nitrogen, sulfur, and complex organic molecules (COMs) was analyzed. We found that an ice-depth dimension, measured, e.g., in monolayers, is essential for modeling of chemistry in interstellar ices. Particularly, the H2O:CO:CO2:N2:NH3 ice abundance ratio regulates the production and destruction of minor species. It is suggested that photodesorption during the core-collapse period is responsible for the high abundance of interstellar H2O2 and O2H and other species synthesized on the surface. The calculated abundances of COMs in ice were compared to observed gas-phase values. Smaller activation barriers for CO and H2CO hydrogenation may help explain the production of a number of COMs. The observed abundance of methyl formate HCOOCH3 could be reproduced with a 1 kyr, 20 K temperature spike. Possible desorption mechanisms, relevant for COMs, are gas turbulence (ice exposure to interstellar photons) or a weak shock within the cloud core (grain collisions). To reproduce the observed COM abundances with the present 0D model, 1%–10% of ice mass needs to be sublimated. We estimate that the lifetime for starless cores likely does not exceed 1 Myr. Taurus cores are likely to be younger than their counterparts in most other clouds.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Low-mass interstellar molecular clouds, such as the Bok globules, are a model case for cloud structure and star formation studies, thanks to their relative simplicity and proximity (Hartmann 2009). For astrochemistry, stable starless cores within these clouds are particularly interesting. They are sufficiently dense for a significant mass of ice to accumulate onto the grains, supposedly not disturbed by gravitational contraction motions. Thus, modeling of surface chemistry can be simplified to calculations that consider constant yet realistic physical conditions.

The aim of the present paper is to study the evolution of the chemical composition of ices in starless cores with physical conditions that are relevant to actual cores. The modeling of surface chemistry in starless cores has a long history. Important papers that have presented significant advances in the field include Watson & Salpeter (1972), Allen & Robinson (1977), Pickles & Williams (1977a, 1977b), Tielens & Hagen (1982), D'Hendecourt et al. (1985), Brown & Millar (1989), Brown (1990), and Hasegawa et al. (1992). A model for small, diffuse clouds (not the starless core) with surface chemistry has been presented by Turner (1998). Such two-phase models with constant physical parameters have also been used in recent papers (e.g., Du et al. 2012; Vasyunin & Herbst 2013b).

Starless cores with physical conditions derived from actual observations have also been modeled (Du et al. 2012; Lippok et al. 2013; Maret et al. 2013; Awad et al. 2014). Recent advances in the field of interstellar ice chemistry in combination with available observational data (e.g., Launhardt et al. 2010) make the modeling of ices in quiescent cores an interesting proposal. These advances include ices described as consisting of layers, with only the surface monolayer (ML) being avalaible for reactions (Hasegawa & Herbst 1993b; Bergin & Langer 1997; Cuppen & Herbst 2007; Garrod & Pauly 2011; Taquet et al. 2012), photoprocessing of subsurface ices (Kalvāns & Shmeld 2010; Garrod 2013; Chang & Herbst 2014), and a realistic model of ice molecule formation.

An important aspect in interstellar ice models is a proper description of surface reactions in a contracting molecular core. Notably, this includes the surface synthesis of CO2 (Ruffle & Herbst 2001; Garrod & Pauly 2011). Brown et al. (1988) and Aikawa et al. (2001) show that the contraction phase is particularly important for a correct representation of the ice formation epoch. Recent dense core models that focus on surface chemistry often do not consider the core contraction phase (Du et al. 2012; Kalvāns & Shmeld 2013; Vasyunin & Herbst 2013b; Chang & Herbst 2014; Reboussin et al. 2014). The dependence of ice composition on the core density and interstellar extinction has been shown by a number of earlier studies (e.g., Rawlings et al. 1992; Viti & Williams 1999; Green et al. 2001).

We studied four starless cores (CB 17—SMM1, CB 26—SMM, CB 27—SMM, and B68) with chemical kinetics modeling. They were treated as clumps of gas and dust with an increased density toward the center, surrounded by a more diffuse cloud. The cores undergo an initial contraction phase, until they reach a state with constant (observed) physical conditions (Section 2.1). A detailed surface and subsurface ice chemistry model has been used (Section 2.2), whose application to starless cores is the main novelty of this paper.

2. METHODS

The model "Alchemic-Venta" was employed in this research, taken from previous works (Kalvāns 2015a, 2015b, hereafter Papers I and II, respectively). It is based on the "ALCHEMIC" code (Semenov et al. 2010), with details explained below. Physical parameters in the model were adjusted for a selection of non-star-forming cores. Papers I and II investigate collapsing prestellar cores, while the present paper focuses on stable, non-star-forming cores.

2.1. Physical Model

2.1.1. Cloud Core Conditions

We investigate ice chemistry of four starless cores with their central density ${n}_{{\rm H}}=n({\rm H})+2n({{\rm H}}_{2})$, in cm−3, and interstellar extinction AV, in mag, derived from observations. Only the core center part was considered (i.e., 0D model), where ${n}_{{\rm H}}$ and AV are at their highest. Each simulation consists of a collapse Phase 1, when the cloud density increases, and a quiescent Phase 2 with constant physical conditions.

A single-point 0D model was chosen because the ices in the center of each core most likely do exist, despite a number of uncertain parameters. These include the efficiency of non-thermal desorption mechanisms and turbulent motions in starless cores, as observed by Redman et al. (2006), Levshakov et al. (2014), and Steinacker et al. (2014). Turbulence was not considered in the present study. Also, ices in the core center likely have experienced the full range of physical conditions of the core since its formation. The simple 0D model puts certain constraints for the interpretation of results.

Cores with a Plummer-like density profile (Plummer 1911; Whitworth & Ward-Thompson 2001), embedded in a surrounding cloud with a fixed column density ${N}_{\mathrm{out}}$, were modeled. The total column density to the core center can be calculated with

Equation (1
)

where ${n}_{{\rm H},0}$ is the (maximum) density at the center of the core, r0 is the radius of the central density plateau, η is the power-law slope at large radii, and r1 is the radius of the core. The chemical modeling was done only for the center of the core, where r = 0 and the density profile of the core is necessary only for the calculation of NH at its center. Table 1 summarizes the parameters ${n}_{{\rm H},0}$, r0, r1, and ${N}_{\mathrm{out}}$. These are unique for each core and were derived from Lippok et al. (2013) for the CB 17, CB 26, CB 27, and B68 cores and from Schmalzl et al. (2014) for the CB 17 core. The two "versions" of CB 17 are hereafter denoted CB 17 S and CB 17 L, for data from Schmalzl et al. (2014) and Lippok et al. (2013), respectively. This means that there are a total of five models (CB 17 S, CB 17 L, CB 26, CB 27, and B68) of four cores considered.

Table 1.  The Constant Physical Parameters of Modeled Starless Cores during Their Stable Phase 2

  Observational Data Calculated Data
 
Core ${n}_{{\rm H},0}$, cm−3 r0, AU r1, AU η ${N}_{\mathrm{out}}$, cm−2 Reference AV, mag T, K t1, kyr
CB 17 S—SMM1 2.3E+5 9.5E+3 3.0E+4 4.9 0 Sa 13.8 8.2 1465
CB 17 L—SMM 1.3E+5 1.1E+4 3.0E+4 5.0 3.6E+20 Lb 9.1 9.2 1451
CB 26—SMM 8.7E+4 8.0E+3 4.0E+4 3.0 2.2E+20 L 6.5 11.1 1437
CB 27—SMM 1.3E+5 1.3E+4 6.6E+4 6.0 3.1E+20 L 9.5 8.9 1451
B68—SMM 4.0E+5 7.0E+3 2.7E+4 5.0 3.0E+20 L 17.6 7.9 1482

Note. Assumed physical parameters: ${n}_{{\rm H},0}$—density at the center of the core; r0—radius of the central density plateau; η—power-law slope for density; r1—radius of the region (core) in consideration; ${N}_{\mathrm{out}}$—column density of the surrounding cloud. Calculated parameter t1 is the length of the cloud core-collapse Phase 1.

aSchmalzl et al. (2014). bLippok et al. (2013).

Download table as:  ASCIITypeset image

Table 1 data are relevant to cores in the stable Phase 2, when the contraction in Phase 1 has ended and the physical parameters of the cores are unchanging. This phase has an assumed integration time of ${t}_{2}\;\gt $ 2 Myr for each modeled core. Prior to the stable phase, a core formation phase was considered. The formation of starless cores has been described before as a free-fall (e.g., Lee et al. 2004) or "delayed free-fall" (e.g., Taquet et al. 2014). A "static," step-like contraction was employed by these authors.

In the present study, the formation of the core was described as a delayed free-fall process. The approach presented by Brown et al. (1988), Nejad et al. (1990), and Rawlings et al. (1992) was used. A Plummer-like core with an initial peak density of ${n}_{{\rm H},\mathrm{ini}}=3000$ cm−3 contracts to the final density ${n}_{{\rm H},0}$ in a core formation time t1 (i.e., length of Phase 1; see Table 1). The parameter B in Equation (1) of Rawlings et al. (1992), which modifies the rate of the contraction, was taken to be 2/3. This means that the collapse to the final densities, indicated in Table 1, occurs in 1.4–1.5 Myr. Figure 1 shows an example of the evolution of physical conditions for a core (B68) during the contraction phase, up to 1.5 Myr.

Figure 1.

Figure 1. Density of hydrogen atoms, interstellar extinction, and temperature of the B68 SMM cloud core—an example of the evolution of physical conditions during the contraction Phase 1. After the contraction phase, these parameters were allowed to remain unchanged for a total integration time of 3.5 Myr.

Standard image High-resolution image

A value B = 0.7 has been used by Nejad et al. (1990), Rawlings et al. (1992), and Ruffle et al. (1999) for prestellar core models. We take $B=2/3$ as a somewhat lower value for starless cores. The exact value of B has only a limited effect on ice composition, the main subject of the present study. This is because ice accumulation was calibrated by modifying the poorly known efficiency of indirect reactive desorption to match the observed proportions of H2O, CO, and CO2 (Section 2.1.4).

For the core formation epoch, r0 was kept proportional to ${n}_{{\rm H},0}^{-1/2}$ (Keto & Caselli 2010; Taquet et al. 2014). The Plummer-like density profile of the contracting core was retained, while the density increases from ${n}_{{\rm H},\mathrm{ini}}$ to ${n}_{{\rm H},0}$. The value of r1, which determines the physical size of the region in consideration, was assumed constant. This means that the mass of the core (i.e., the number of H atoms within a sphere of radius r1) increases during the formation phase. ${N}_{\mathrm{out}}$ was retained constant, too.

The total integration time ${t}_{1}+{t}_{2}$ was taken to be 3.5 Myr, and the results were analyzed for an interval of 1.0–2.3 Myr (see Section 3.3). In context with ice formation, the full time consists of three periods: (1) cloud with ${n}_{{\rm H},0}\approx 3\times {10}^{3}$ cm−3 and no ice on the grains; (2) ice accumulation up to a maximum thickness in a rapidly contracting core and up to 200 kyr into the quiescent phase; and (3) slow processing of ice by cosmic-ray-induced photons after the freeze-out has ceased. The start of period 2 is marked with the appearance of first subsurface ice species at densities 5 × 103 to 1 × 104 cm−3. During the last period (3), the gas and the surface are in an approximate adsorption–desorption equilibrium. The nominal ice thickness decreases during period 3.

Ice photoprocessing leads to a declining amount of CO in the icy mantles. Because of this, quiescent Phase 2 integration times longer than 2 Myr were not considered. For longer times, the H2O:CO:CO2 ice abundance ratio is no longer supported by observations of interstellar ices (Gibb et al. 2004; Whittet et al. 2007; Boogert et al. 2011; Öberg et al. 2011). Mantle growth and the abundances of major ice species are considered in Section 3.

Gas and dust were assumed to be in thermal equilibrium with temperature derived from the interstellar extinction AV, in line with Garrod & Pauly (2011). In other words, it was assumed that the cores are heated externally by the interstellar radiation field. AV was calculated with

Equation (2
)

Table 1 shows the calculated values for AV and T for each of the modeled cores in the stable phase.

Table 2 shows the chemical abundances used at the start for all simulations. They are based on the abundances derived by Garrod & Herbst (2006) and Wakelam & Herbst (2008). The rate of hydrogen ionization by cosmic rays was taken to be $1.3\times {10}^{-17}$ s−1. The flux of cosmic-ray-induced photons and interstellar photons was taken to be 4875 and 108 cm−2 s−1, respectively, with a grain albedo of 0.5.

Table 2.  Initial Abundances of Chemical Species

Species ${n}_{i}/{n}_{{\rm H}}$
H2 0.4995
H 1.00E-03
He 9.00E-02
C 1.40E-04
N 7.50E-05
O 3.20E-04
Na 2.00E-08
Mg 2.55E-06
Si 1.95E-06
P 7.63E-08
S 1.50E-06
Cl 1.40E-08
Fe 7.40E-07

Download table as:  ASCIITypeset image

The self- and mutual shielding of H2, CO, and N2 was included with the use of the tabulated shielding functions (Lee et al. 1996; Li et al. 2013). Shielding was also attributed to molecules in ice. It is important to note that the surrounding molecular cloud was included in these calculations by assuming an external H2 column density of ${N}_{\mathrm{out}}/2$. This was not included in Papers I and II, which used a physical model built on similar principles. Such a more accurate approach here is of particular importance, because more plausible values for indirect reactive desorption efficiency can now be derived (Section 2.1.4). The synthesis of hydrogen on the grain surface was considered by both Langmiur–Hinshelwood and Eley–Rideal mechanisms, although the latter is of little importance.

2.1.2. Dust Grain Properties

The grains were assumed to consist of two components. The first is inert, solid nuclei with a radius of $a=0.1\;\mu {\rm m}$ and an abundance $1.32\times {10}^{-12}$ relative to H atoms. In dark and dense conditions (${A}_{V}\gt 1.9$ mag), the second component, interstellar ice, begins to accumulate onto grain surfaces in significant amounts. The increase of grain size is calculated self-consistently, assuming a thickness of 350 pm for each ML, which is multiplied by the number of ice MLs to obtain ice thickness b. The number of adsorption sites per ML was retained constant at $1.5\times {10}^{6}$. Changes in this number likely do not influence the agreement of model results with observations (Acharyya et al. 2011; Taquet et al. 2012) and were not considered in the model. Dust grain albedo was assumed to be 0.5.

The ice layer on the grains was described with the sublayer approach, developed in Paper II. This means that a fully formed ice (accumulation from gas has ceased) consists of a surface layer (1–2 MLs) and three subsurface layers—"sublayers"—of approximately equal thickness. A sublayer nominally may contain up to several tens of MLs. Two or more sublayers in a model allow us to consider the depth-dependent composition of the ice mantles. Three sublayers were considered in the present study, similarly to Paper II. For reference, they were numbered with the first being the shallowest and the third being the sublayer adjacent to the grain nucleus. The transition from the surface to the first sublayer and, subsequently, to the second and third sublayers has been harmonized with the accretion rate, in line with Papers I and II.

2.1.3. Ice Physical Description

The molecules are bound to the surface or into the mantle with their adsorption (ED) or absorption energies (EB), respectively. ED is the "basic value," given in the reaction network, from which all the other energies were calculated. The binding energies, used for calculating the diffusion rates, for surface (${E}_{b,S}$) and mantle (${E}_{b,M}$) species were assumed to be 0.40ED and 0.40EB, respectively (Garrod & Pauly 2011). In Papers I and II it was argued that ${E}_{b,M}$ has to be higher, or similar to ED; otherwise, the movement of molecules in ice bulk would be faster than their evaporation from the surface, which fits the description of a liquid, not a solid. In line with the earlier works, it was assumed that ${E}_{B}=3{E}_{D}$, i.e., that typically ${E}_{b,M}=1.2{E}_{D}$. For CO and CO2 molecules, the ratios ${E}_{b,S}/{E}_{D}={E}_{b,M}/{E}_{B}$ were taken to be 0.31 and 0.39, respectively (Karssemeijer & Cuppen 2014).

In dark cores, the grains may be sufficiently cold to have a significant proportion of their surface covered by molecular hydrogen. H2 provides a much weaker binding than water-ice. The adsorption energy for H and H2 was calculated according to Garrod & Pauly (2011). For other species, ED was not affected by the surface coverage of H2 because of the large difference between the hopping rate of H2 and vibration frequency of heavy molecules (Paper II).

Accretion of neutral gaseous species onto the grains with a radius $a+b$ was calculated with sticking probabilities of 0.33 for light species (H, H2) and 1.0 for all other species (Kalvāns & Shmeld 2010). A total of six desorption mechanisms were considered. The rate coefficients for evaporation and cosmic-ray-induced whole-grain heating (Hasegawa & Herbst 1993a) were calculated for each molecule as a thermal desorption at the equilibrium temperature and 70 K, respectively. Desorption by interstellar and cosmic-ray-induced photons was considered with uniform yields for all species, 0.003 and 0.002, respectively. These values were chosen to be in agreement with recent photodesorption experiments with interstellar ice analogs, as discussed in Paper II. Photodesorption from subsurface mantle species was permitted, if the number of overlying MLs does not exceed two. The yield for subsurface photodesorption was reduced by a factor of $1/3$. This approach was developed to be in line with the results of Andersson & van Dishoeck (2008).

A reaction-specific reactive desorption was used, with the ${\alpha }_{\mathrm{RRK}}$ parameter set to 0.03 (Paper II). This is the ratio of the surface-molecule bond frequency to the frequency at which energy is lost to the grain surface (Garrod et al. 2007).

2.1.4. Indirect Reactive Desorption

Desorption induced by the energy released from the H+H exothermic surface reaction probably determines the relative proportions of major ice components H2O, CO, and CO2. In order to reproduce observations of interstellar ices, a sequence $\mathrm{CO}\gt {\mathrm{CO}}_{2}\gg {{\rm H}}_{2}{\rm O}$ must be observed for indirect reactive desorption efficiency ${f}_{{{\rm H}}_{2}\mathrm{fd}}$ (Paper II). Without indirect reactive desorption, the calculated abundance of solid carbon species in collapsing cores can be overestimated, when compared to the observational results provided by Gibb et al. (2004), Whittet et al. (2007), and Öberg et al. (2011). For example, realistic CO and CO2 relative abundances can be achieved, if the ice contains a substantial amount of CH4, H2CO, and (or) CH3OH (Garrod & Pauly 2011; Taquet et al. 2014). Of the latter three species, only methanol has been occasionally observed in interstellar ices (Boogert et al. 2011).

The parameter ${f}_{{{\rm H}}_{2}\mathrm{fd}}$ can be described as the number of desorbed molecules per H2 formed on a grain. Exact values for the efficiency are unknown, and given the averaging and uncertainties in astrochemical numerical simulations, it is reasonable to adjust ${f}_{{{\rm H}}_{2}\mathrm{fd}}$ so that it produces an approximate fit with observations of starless cores. Such an approach can be used until a more precise value of ${f}_{{{\rm H}}_{2}\mathrm{fd}}$ is found in experiments or, for example, quantum chemistry calculations.

In order to derive ${f}_{{{\rm H}}_{2}\mathrm{fd}}$ for the present study, we employ a similar method to Paper II. A prestellar core in free-fall gravitational collapse was considered in that study. For the present research, the B68 core model (Table 1) was used. It has the highest final density and final AV among all the starless cores in consideration. This allows the comparison of observed/modeled CO:H2O and CO2:H2O ice abundance ratios for five AV values from the data set of Whittet et al. (2007). An additional observational constraint for the model is the abundance of H2O, CO, and CO2 ices at their respective threshold extinctions ${A}_{\mathrm{th}}$.

The above means that, while the B68 cloud core is currently stable (Bergin et al. 2006; Redman et al. 2006), we assume that it formed in a delayed free-fall process, which ceased when the core reached its present density. This assumption has been attributed to all simulated cores in this study. While it might not be entirely correct, it is probably just as useful as other realistic assumptions, because of the poorly constrained value of ${f}_{{{\rm H}}_{2}\mathrm{fd}}$, which has to be calibrated, regardless of the exact core formation path.

Table 3 summarizes the observational and best-fit B68 model results. The desorption efficiency ${f}_{{{\rm H}}_{2}\mathrm{fd}}$ was calculated according to the empirical approach of Paper II:

Table 3.  Comparison of Observed and Calculated (B68 Core) CO and CO2 Ice Abundances, Relative to Water

    Observations B68 Model  
Source ID AV
tb, kyr
043728.2 + 261024 6.3 ± 1.5 . . . 25.0 1.8 13.8 1417
042324.6 + 250009 10.0 ± 0.5 20.2 19.1 7.7 13.0 1451
043325.9 + 261534 11.7 ± 0.5 12.0 16.7 12.5 13.9 1459
043926.9 + 255259 15.3 ± 0.5 27.3 16.7 30.2 16.5 1471
042630.7 + 243637 17.8 ± 1.5 45.1 18.3 43.2 18.2 1476
Threshold Ice Abundances   H2O, ML CO2, ML CO, ML  
${A}_{\mathrm{th}}$ H2O 3.2 ± 0.1 3.5 1332
${A}_{\mathrm{th}}$ CO2 4.3 ± 1.0 1.5 1375
${A}_{\mathrm{th}}$ COc 6.8 ± 1.6 0.6 1424
${A}_{\mathrm{th}}$ COd 8.1 1.3 1437

Notes. Observational data from Whittet et al. (2007), unless noted otherwise.

aAbundance ratio for icy species. bIntegration time, corresponding to the AV value. cWhittet et al. (2001). dPaper II.

Download table as:  ASCIITypeset image

Equation (3)

where $Q=9.5\times {10}^{-5}$ and ${Q}_{1}={10}^{3}$. The adopted best-fit values of Q and O1 are different from those in Paper II. It is possible that the values obtained here are more reliable than those of Paper II because of the more advanced and realistic physical model (Section 2.1.1).

The observational values for CO:H2O and CO2:H2O ice abundance ratios were taken from Whittet et al. (2007). While newer data sets are available (Boogert et al. 2011; Öberg et al. 2011), Whittet et al. (2007) provide the AV for each background star, which enables us to tie the observations to specific time points of the simulations with the corresponding AV (see Paper II for more on this approach).

2.2. Chemical Model

2.2.1. Reaction Network

Several changes have been introduced into the gas-grain network, which was acquired from Laas et al. (2011). The four new gas-phase reactions of Vasyunin & Herbst (2013b) were added. All these reactions involve the methoxy radical CH3O. Because these authors did not discern between the methoxy radical and its isomer, the hydroxymethyl radical CH2OH, four complementary reactions were added for the latter species (1–4 in Table 4).

Table 4.  Changes in Reactions Respective to the Original Network by Laas et al. (2011)

No. New Gas Phase Reactionsa α β γ Reference
1 ${\mathrm{CH}}_{2}\mathrm{OH}+{\rm H}\;\longrightarrow \;{\mathrm{CH}}_{3}+\mathrm{OH}$ 1.6E-10 0 0 NIST
2 ${\mathrm{CH}}_{2}\mathrm{OH}+{\rm O}\;\longrightarrow \;{{\rm H}}_{2}\mathrm{CO}+\mathrm{OH}$ 1.5E-10 0 0 NIST
3 ${\mathrm{CH}}_{2}\mathrm{OH}+{\mathrm{CH}}_{3}\;\longrightarrow \;{{\rm H}}_{2}\mathrm{CO}+{\mathrm{CH}}_{4}$ 1.4E-10 0 0 NIST
4 ${\mathrm{CH}}_{2}\mathrm{OH}+{\mathrm{CH}}_{3}\;\longrightarrow \;{{\rm C}}_{2}{{\rm H}}_{5}\mathrm{OH}$ 1.0E-15 −3.00 0 Vasyunin & Herbst (2013b)b
No. New Surface (ice) Reactions EA, K Reference
5 ${\rm H}+\mathrm{HCO}\;\longrightarrow \;\mathrm{CO}+{{\rm H}}_{2}$ 0 This work
6 $\mathrm{OH}+\mathrm{HCO}\;\longrightarrow \;\mathrm{CO}+{{\rm H}}_{2}{\rm O}$ 0 This work
7 ${\mathrm{NH}}_{2}+\mathrm{HCO}\;\longrightarrow \;\mathrm{CO}+{\mathrm{NH}}_{3}$ 0 This work
8 ${{\rm H}}_{2}+\mathrm{HCO}\;\longrightarrow \;{\mathrm{CH}}_{3}{\rm O}$ 2100c This work
9 $\mathrm{OH}+\mathrm{SO}\;\longrightarrow \;{\mathrm{SO}}_{2}+{\rm H}$ 0 Gas phase
10 ${\rm O}+\mathrm{HS}\;\longrightarrow \;{\rm S}+\mathrm{OH}$ 956 Gas phase
11 ${\rm O}+\mathrm{HS}\;\longrightarrow \;\mathrm{SO}+{\rm H}$ 0 Gas phase
12 ${\rm H}+\mathrm{CL}\;\longrightarrow \;\mathrm{HCL}$ 0 Gas phase
No. Changed Surface (ice) Reactions EA, K Reference
13 $\mathrm{CO}+{\rm O}\;\longrightarrow \;{\mathrm{CO}}_{2}$ 290 Roser et al. (2001)
14 ${{\rm O}}_{2}+{\rm H}\;\longrightarrow \;{{\rm O}}_{2}{\rm H}$ 600 Du et al. (2012)
15 ${\rm H}+{{\rm H}}_{2}\mathrm{CO}\;\longrightarrow \;{\mathrm{CH}}_{3}{\rm O}$ 2100 Woon (2002)
16 ${{\rm O}}_{3}+{\rm H}\;\longrightarrow \;{{\rm O}}_{2}+\mathrm{OH}$ 0 Mokrane et al. (2009)

Notes.

aIn addition to the four reactions from Vasyunin & Herbst (2013b). Reaction rate coefficient $k=\alpha {(T/300)}^{\beta }\mathrm{exp}(-\gamma /T)$. bData adopted from the ${\mathrm{CH}}_{3}{\rm O}+{\mathrm{CH}}_{3}$ reaction. cEstimate.

Download table as:  ASCIITypeset image

The results of Paper II show that formic acid HCOOH, formaldehyde H2CO, and formamide NH2CHO can be overproduced in subsurface mantle conditions with the surface reaction set of Laas et al. (2011), which has been adopted from Garrod et al. (2008). In an attempt to reduce their abundance, grain surface reactions 5–8 in Table 4 have been added to the network. They tend to reduce the production efficiency of HCOOH, H2CO, and NH2CHO because an alternative outcome is now possible for their reactants.

Reactions 5–7 were assumed to be barrierless. Reaction 8 has an assumed activation barrier EA = 2100 K, estimated from reactions of H2 with compounds that have a similar structure to HCO. The barrier of the latter reaction is so high that it has little effect on the results of the current cold core model.

Reactions 9–11 were added to more adequately reproduce the chemistry of sulfur oxides, with activation barriers taken from similar gas-phase reactions. Finally, reaction 12 was added to avoid most of the chlorine atoms ending up in solid ClO, which is probably unrealistic, given the availability of atomic hydrogen, irradiation, and an aqueous medium on grain surfaces. The activation barriers for four additional surface reactions—13–16 in Table 4—were also changed, in line with available data.

2.2.2. Surface and Bulk-ice Chemistry

The rate for reactions on grain surface was calculated according to the reaction–diffusion competition approach of Garrod & Pauly (2011). The modified rate equations method has been applied (see Paper II). As explained in Section 2.1.3, part of the surface reaction products desorb into the gas. Reaction barriers are overcome either thermally or by quantum tunneling, whichever is faster (Garrod et al. 2008).

The rate of binary reactions in bulk ice was calculated for reactants that are immobile for most of the time (Papers I and II). It was assumed that each molecule vibrates in a lattice cell with 10 neighboring molecules. Reactants have to overcome a certain energy barrier ${E}_{\mathrm{prox}}=0.1{E}_{D}$ in order to achieve sufficient proximity for a reaction to occur. Molecule diffusion in ice and reaction–diffusion competition have been taken into account in a manner similar to that for surface reactions. I refer the reader to Papers I and II for a more detailed justification and formalism for the calculation of reaction rate coefficients.

Surface and bulk-ice molecules are subjected to photodissociation by interstellar and cosmic-ray-induced photons. For surface species, the dissociation rate has been taken to be equal to that of gas-phase species. For molecules within the mantle, the photon flux is attenuated by overlying MLs. Each ML has an assumed absorption probability of 0.007 (Andersson & van Dishoeck 2008). In order to calculate the photon flux, respective to each sublayer, the value relevant to the middle ML of that sublayer was used (see Paper II).

3. RESULTS: ICE FORMATION AND MAJOR COMPONENTS

To avoid repetition in the studies of five roughly similar models of objects of the same class, the modeling results are not reviewed for each case separately. Instead, we focus on several astrochemical problems, relevant for ices in starless cores. The use of five models with physical conditions derived from observations instead of an abstract model adds to the credibility of the research. We also have to keep in mind that the 0D model adequately represents only the center part of each core. The abundances values used here are calculated with N(X)/N(H), i.e., relative to total hydrogen. This means that observationally detected abundances that are given relative to H2 were divided by 2 for comparison with calculation results.

3.1. The Evolution of the Ice Layer

The ice layer is basically nonexistent (around 0.1ML) when ${A}_{V}\lt 1.9$ mag, which means that the calculation results (abundances of species) are irrelevant for ice chemistry if $t\leqslant 1$ Myr. Because of this, we interpret results for integration times $t\gt 1$ Myr only. Ice formation observes the rules outlined in Paper II. As the cloud contracts and AV increases over 2 mag, ice molecules begin to accumulate on grain surfaces, forming a mixture of H2O, CO2, and NH3. The ice is dominated by interstellar photons, which means that all surface CO is oxidized to CO2 via the sequence

Equation (4)

probably the most important chemical transformation in bulk ice. When the cloud becomes darker, the flux of interstellar photons is diminished, rapid ice photoprocessing becomes impossible, and CO begins to accumulate at ${A}_{V}\gt 5$. The accretion of CO is faster than the surface synthesis of CO2, and CO becomes the second most abundant ice molecule when ${A}_{V}\gt 12$ mag.

Figure 2 shows the nominal thickness of ice and the abundances in MLs for a few most important species in each sublayer for all five starless cores. Ice thickness peaks shortly after the end of the core-collapse stage and then slowly decreases. The reason for this decrease is the conversion of simple ice species into more complex ones via photoprocessing. Most importantly, sequence (4) converts water and carbon monoxide into carbon dioxide. Thus, the number of molecules per grain and the nominal ice thickness are reduced. This can be seen in Figure 3. A significant amount of hydrogen is released in the process. H2 partially accumulates in the mantle when $T\lt 10\;{\rm K}$. The inner sublayer 3 is the least affected because it has a lower proportion of CO and it is better shielded than the above layers. These results are different from those of Chang & Herbst (2014) thanks to the gradual increase of gas density and AV during Phase 1, which ensures an ice composition that is much more differentiated between layers. Such a chemical differentiation (H2O in inner and CO in outer layers) has been confirmed by models that consider ice formation in collapsing cores with ML accuracy (Garrod & Pauly 2011; Taquet et al. 2014).

Figure 2.

Figure 2. Calculated ice thickness and abundance in MLs for major species in the sublayers for the modeled cloud cores. The number in parentheses indicates the number of the respective sublayer, while "Other" stands for all other icy species in that sublayer. The vertical gray line indicates t50 for each model.

Standard image High-resolution image
Figure 3.

Figure 3. Calculated abundances, relative to H atom number density, of major species in ice for the five modeled cores.

Standard image High-resolution image

CB 26 is the smallest of the five cores; it has the lowest final AV and density values (Table 1). This means that molecules on grains deposit at a lower rate and have undergone significant subsurface photoprocessing even before the freeze-out has ended. We understand the "freeze-out epoch" as the interval between the time of formation for the first sublayer (${t}_{\mathrm{sub}}$) and the time when maximum ice thickness is reached (${t}_{\mathrm{max}}$, as specified in Table 5). Parameter t50 is the point in time when the CO2:H2O ratio in ice has grown to 50%, and for CB 26 it occurs even before ${t}_{\mathrm{max}}$.

Table 5.  Calculated Ice Composition Parameters in the Starless Core Models

    ${t}_{\mathrm{sub}}$, Max. Ice Thickness t50, Ice at 3.5 Myr
Core AV Myr ML ${t}_{\mathrm{max}}$, Myr
Myr ML
CB 17 S 13.8 1.25 170 1.57 81 23 2.31 147 68 81
CB 17 L 9.1 1.21 164 1.62 74 35 1.92 141 61 107
CB 26 6.5 1.21 153 1.68 69 59 1.57 133 48 150
CB 27 9.5 1.06 158 1.60 69 51 1.58 131 51 175
B68 17.6 1.15 171 1.52 78 25 2.12 146 63 90

Note. ${t}_{\mathrm{sub}}$—time for ice thickness to reach 1ML; ${t}_{\mathrm{max}}$—time of maximum ice thickness; t50—time for the CO2:H2O ice abundance ratio to reach 50%.

Download table as:  ASCIITypeset image

On the other hand, CB 27 is the largest core considered. Paradoxically, this results in a similar ice composition to that of CB 26. The AV threshold for accumulation of a significant mass of ice on the grains (>1ML) is close to 1.9 mag. This occurs at a density of ≈5 × 103 cm−3 for CB 27 and in the range $(8-9)\times {10}^{3}$ cm−3 for the other four cores. The result is that the first ice sublayers in CB 27 appear some 150 kyr earlier than in other cores. The formation of sublayer 1 in CB 27 is complete some 20 kyr earlier than for the other cores. This means that sublayer 1 has formed in a relatively diffuse environment, which is still largely dominated by interstellar photons, where water photodissociation products combine with CO on the surface to produce an ice that contains a large proportion of CO2.

Because indirect reactive desorption, which largely governs ice composition, has been calibrated for the B68 core (Section 2.1.4), the explained mechanisms may lead to different results, if ${f}_{{{\rm H}}_{2}\mathrm{fd}}$ is different. The important thing is that both a diffuse medium and early ice formation may lead to a significantly photoprocessed ice. The first is characteristic for low-mass molecular cloud cores, while the second is more likely for relatively massive cores.

Table 5 summarizes data for each model at two important points in time: when ice thickness is at its maximum, and at the end of the simulation run. The thickest ice layer is attained in CB 17 S and B68 models. High ice thickness for B68 can be expected because of its high extinction and final density. For CB 17 S, neither its density nor extinction is the highest. A thick ice layer for this core arises because ${N}_{\mathrm{out}}$ for this model is 0, which means that interstellar photons are attenuated to a lesser extent and the ice layer begins to form at a later, denser stage of the collapsing core. In turn, ice accumulates rapidly and undergoes less intense photoprocessing, and a smaller number of CO and H2O molecules are converted to CO2. This results in higher nominal ice thickness. This conclusion is supported by the abundance of carbon oxides—a higher proportion of CO2 means a nominally thinner ice (Table 5).

The above analysis confirms that extinction largely regulates the formation of the ice layer on the grains because AV determines the number of interstellar photons that reach the grains (Watson & Salpeter 1972). Interstellar extinction also affects the abundance of radical species on surfaces and in the gas, which determines the efficiency of direct and indirect reactive desorption. It can be safely said that the formation history of a starless or prestellar core determines its ice composition.

3.2. General Ice Chemistry

Aside from water and carbon oxides, the nitrogen species NH3 and N2 also can be considered major ice constituents. Their abundance is little affected by ice photoprocessing. Other ice species whose abundance may exceed 1% relative to water for at least one of the modeled cores include H2, O2, H2O2, O2H, and H2S.

Figure 2 shows ice evolution up to 3.5 Myr, the full length of the simulation runs. Although the modeled cores have different physical parameters, ice sublayers share a similar composition in all cases. The inner sublayer 3 contains a large amount of water, in addition to CO2 and NH3. All CO molecules here are quickly consumed, which makes chemical processes in sublayer 3 somewhat different from those in the above layers. The lack of CO and its daughter atomic C means that O, OH, N, NH, S, and other radicals are more available.

Sublayer 2 initially forms as a ≈1:1 mixture of H2O and CO, with additions of CO2 and N2. As time progresses, CO and H2O combine, and the sublayer becomes increasingly dominated by CO2. The outer sublayer 1 largely consists of CO, with an admixture of H2O, CO2, and N2 at roughly similar proportions. CO remains the main species in this layer, although the proportion of CO2 is steadily growing. Sublayer composition may profoundly affect the chemistry of minor species because the main reacting species are photodissociation fragments of major molecules.

3.3. Comparison with Observations and Age Limits

The initial chemical composition of ices is largely regulated by indirect reactive desorption, calibrated with the B68 model (Section 2.1.4). Table 3 shows that carbon oxides are somewhat depleted during the earlier stages in cloud contraction, compared to the observational data by Whittet et al. (2007). We found it impossible to remove this discrepancy with adjustments in the efficiency of indirect reactive desorption, Equation (3). This depletion may arise, for example, if the calculated abundance of atomic H is higher by a factor of two than H abundance in the real core. Another possible cause can be changes in surface properties as proportionally more and more CO molecules stick to the grains during the ice formation epoch. We will investigate the latter suggestion in a separate paper.

Figure 4 shows that molecules, typically formed on grain surfaces or dense gas (e.g., H2O, CO2, NH3, and O2), are also abundant in the gas phase during or immediately after the core contraction stage, when density is in excess of ≈104 cm−3. Gas and surface molecules are actively interchanged during the freeze-out epoch. Such gas-phase abundance peaks in a layer around a collapsing core have been found in models of prestellar cores (Garrod & Herbst 2006; Vasyunin & Herbst 2013a; Paper II) and observations of water vapor in the prestellar core L1544 (Caselli et al. 2012). The present model predicts that such peaks are likely to be shorter and higher thanks to more efficient desorption mechanisms. Desorption retains a substantial amount of refractory species in the gas phase until very late times in core contraction.

Figure 4.

Figure 4. Calculated abundance of major gas-phase molecules, relative to hydrogen.

Standard image High-resolution image

We can conclude that the Phase 1 modeling results confirm that gas falling into a prestellar or protostellar core is rich with molecules synthesized on the surface. The abundance peaks predicted by the present study are higher than in other similar models, thanks to more efficient desorption mechanisms. The main desorption agents are interstellar photons, as suggested by Dominik et al. (2005).

Subsurface ice processing may place constraints for the maximum age of starless cores. The observed solid CO2:H2O abundance ratio does not exceed 44% (Boogert et al. 2011), although higher values are permitted by upper limits. As a threshold value for the age of the cores, we chose the longest time taken for any of the cores to reach a CO2:H2O ice abundance ratio of 50%, corresponding to a time t50. For the CB 17 S core ${t}_{50}=2.31$ Myr, longer than t50 for other cores (Table 5). A single end time of 2.31 Myr was applied for all simulations. The decision to adopt a single and rather long total integration time for all models leads to the result that for the CB 26 and CB 27 cores CO2:H2O is higher than 50% most of the time. With this approach the model considers a variety of icy environments and allows for an easy comparison between the five simulation results.

The above means that the actual lifetime t2 of a quiescent Phase 2 core with constant physical conditions is below 0.9 Myr for all models. This is significantly less than the initial assumption of ${t}_{2}\;\approx $ 2 Myr for a 3.5 Myr total simulation length. Thus, the modeling results were analyzed for an integration time interval of 1.00–2.31 Myr.

If gas and dust in the center part of a starless core are not effectively mixed with its surroundings, CO2-rich ices should be present in starless cores. Turbulence and mixing of gas in the dense cores are poorly known parameters that may contribute to the destruction of the ices (Boland & de Jong 1982; Williams & Hartquist 1984; Redman et al. 2006; Adams & Shu 2007; Levshakov et al. 2014; Steinacker et al. 2014). Given the uncertainties, we treat the chemistry of CO2-rich ices as a theoretical possibility. Their observational detection or non-detection might provide additional constraints on the physicochemical processes that govern ice formation, chemical processing, and desorption.

4. RESULTS: ELEMENTAL CHEMISTRY

4.1. Oxygen Chemistry in Ice

4.1.1. Chemical Processing of Minor Oxygen Species

Three major ice species—H2O, CO, and CO2—contain copious amounts of the eighth element, and the chemistry of oxygen in ice is closely connected to these molecules. When water and carbon oxides are excluded, the bulk-ice chemistry of oxygen for CB 17 S and B68 is initially dominated by molecular oxygen O2, which is rather quickly converted into hydrogen peroxide H2O2, as shown in Figure 5. For other models, where ice photoprocessing is more intense, H2O2 is the more important species from the beginning. H2O2 is more stable in irradiated interstellar ices than O2, which mainly arrives from the gas phase. The hydroperoxyl radical O2H is the most abundant radical for oxygen chemistry in ice, thanks to its high ${E}_{b}$. It is generated by two mechanisms—H addition to O2 (early times) and the photodissociation of H2O2 (late times). Hydroxyl radical OH begins to accumulate to a limited extent in the inner sublayer 3 only after the CO molecule has been oxidized (Figure 6).

Figure 5.

Figure 5. Calculated abundance of selected oxygen species in ice, relative to hydrogen.

Standard image High-resolution image

Interestingly, CB 26, which has the lowest AV and highest flux of interstellar photons, has also the lowest radical content. This is because the higher temperature in CB 26 (see Table 1) allows a greater radical mobility (especially for atomic H) in the mantle and, thence, reactivity. Low radical content in CB 26 ices is characteristic also for other compound classes (Sections 4.2, 4.3, 5.1).

Because ice in CB 17 S is initially rich in O2 (Figure 5), it has also a higher abundance for H2O2, which forms via reaction 14 (Table 4). Substantial amounts of H2O2 (relative abundance ${X}_{{{\rm H}}_{2}{{\rm O}}_{2}}$ in excess of 10−6) in sublayer 3 for CB 17 S and B68 are able to form from water photodissociation products that combine together after all CO has been oxidized in that sublayer.

Figure 6.

Figure 6. Calculated relative abundance of main O and O–H species in ice mantle sublayers for the CB 17 S and CB 26 models. The latter represents a significantly photoprocessed ice. The irregularities for the CB 17 S surface layer at $t\approx 1.5$ Myr are artifacts.

Standard image High-resolution image

4.1.2. Comparison with Observations

Figure 7 shows that the calculated abundance of gaseous hydrogen peroxide H2O2 can be reasonably high for a short period, up to 9 × 10−10 relative to the total hydrogen atomic number density for CB 17 S. The abundance of its associated O2H radical, however, is lower by three orders of magnitude. Both these species have been observed in interstellar medium with roughly similar abundances of 5 × 10−11 (Bergman et al. 2011; Parise et al. 2012). It has been suggested that reactive desorption is responsible for transferring these species from the surface to the gas (Du et al. 2012). However, reactive desorption may be too inefficient to maintain a significant amount of such heavy species in the gas (Paper II). The comparison of abundances here is indicative, only, because of the limits of the 0D model.

Figure 7.

Figure 7. Calculated abundance of selected oxygen species in the gas phase, relative to hydrogen.

Standard image High-resolution image

Because practically all of H2O2 in ice resides in the sublayers, photodesorption from subsurface ice may play a role in the existence of gas-phase H2O2 and O2H. More often than not, dissociative photodesorption occurs, when molecule fragments are ejected, instead of intact species (Andersson et al. 2006; Andersson & van Dishoeck 2008). This aspect has not been considered in the present model, or in the model by Du et al. (2012). It is likely that photodesorption of H2O2 also induces a high gas-phase abundance of O2H.

Beginning with Goldsmith et al. (2002), molecular oxygen O2 has been repeatedly observed in shocked gas. A few detections have been associated with the cold dense core ρ Ophiuchi A, with inferred abundances in the range $2.5\times {10}^{-7}\;\mathrm{to}\;1.25\times {10}^{-6}$ relative to total hydrogen. Such high abundances likely can be observed only during a short period in the core's evolution (Larsson et al. 2007; Liseau et al. 2010). These results agree with the calculation results shown in Figure 7, where the maximum abundances lie in the range $7\times {10}^{-7}\;\mathrm{to}\;4.5\times {10}^{-6}$. A quantitative comparison is possible in this particular case because of the centrally localized nature of the observed source (Liseau et al. 2010).

4.2. Nitrogen Chemistry in Ice

4.2.1. Chemistry of Important Species

Figures 8 and 9 show that ammonia NH3 and molecular nitrogen N2 are the main nitrogen reservoirs in ice and gas. Nitrogen chemistry is regulated by proportions of N2 and NH3 in the surface and the three sublayers (Section 3.2). The relatively clear division of NH3 (in sublayer 3) and N2 (in sublayers 1 and 2) means that N, NH, and NH2 radicals are not equally available in all sublayers. Interesting examples of relatively abundant minor nitrogen species in ice are formamide NH2CHO (discussed in Section 5), hydrogen cyanide HCN, and isocyanide HNC.

Figure 8.

Figure 8. Calculated abundance for important nitrogen species in ice, relative to hydrogen.

Standard image High-resolution image

The chemistry of HCN and HNC is controlled by the proportions of CO, N2, and NH3 in each sublayer. HNC is more favored by ice chemistry at earlier times than cyanide HCN, while HCN is the more favored isomer during late times. The two isomers form via different pathways. HNC is mainly synthesized in the reaction NH2+C, while HCN is the product of H addition to CN. The latter forms via C+N. This means that HNC is a product of NH3 and CO mixture photoprocessing, while HCN is synthesized in an N2 and CO mixture. The HNC pathway is more efficient because it requires no subsequent hydrogenation. The result is that 60% (for CB 17 S) to 90% (for CB 26) of both isomers is concentrated in the outer sublayer 1, where CO and N2 are highly abundant, and NH3 is available, too. The remaining 10%–40% resides in sublayer 2, while sublayer 3 is very poor on HCN and HNC (<2%), because of a lack of CO there. Because the surface closely interacts with gas-phase CO and N2, HCN is the main isomer for the top ice layer.

4.2.2. Comparison with Observations

Recent observed NH3 gas-phase abundances lie in the range $9\times {10}^{-9}...7.5\times {10}^{-8}$ relative to total H (Ruoskanen et al. 2011; Levshakov et al. 2014). Crapsi et al. (2007) derives a central NH3 abundance of $4.5\times {10}^{-9}$ for the L1544 dense core. These values qualitatively agree with calculation results for cores with ages in the range t = 1.5–1.8 Myr.

Figure 9.

Figure 9. Calculated abundance of important nitrogen species in the gas phase, relative to hydrogen.

Standard image High-resolution image

Figure 10 shows the calculated HNC:HCN abundance ratio of the center of the respective dark cores in gas and solid phases. We note that the observed gas-phase values of this ratio are in the range of 0.54–4.5, with no distinction between YSOs and starless cores (Hirota et al. 1998). The HNC:HCN ratio in ice may exceed unity for all cores, except for CB 26, where lower abundance of CO prevents the buildup of HNC. The highest calculated HNC:HCN ratio is in the CB 27 ice phase, where it reaches 1.75 briefly at 1.54 Myr. Such a peak arises because of effective HNC production in sublayer 1, not observed in other simulations. In turn, this occurs because sublayer 1 in CB 27 is richer in NH3 by a factor of >2 than in other models. Finally, NH3 is more abundant because ices in CB 27 are accumulated at lower extinctions, which means a higher proportion of species that form from free atoms on the surface (H2O, NH3, CH4, etc.).

Figure 10.

Figure 10. Calculated HNC:HCN abundance ratio in gas and ice.

Standard image High-resolution image

4.3. Sulfur Chemistry in Ice

Figures 11 and 12 show that H2S and SO are the most abundant sulfur species in ice and gas. H2S is more abundant than SO for most of the time. An exception is the CB 26 core, where H2S is the only dominant species and SO ice forms with a maximum relative abundance of 5 × 10−8 at 1.56 Myr, only, thanks to surface reactions. Then, SO is quickly consumed by ice mantle reactions because of the higher temperature in the CB 26 core. In the other four cores, the HS radical accumulates and has time to react with atomic O, producing SO. Increased mobility of hydrogen in CB 26 ice does not allow the accumulation of HS, S, and many other radicals. H2S can be expected to be the main sulfur reservoir for photon-dominated ices. Notably, CB 26 still has a maximum SO gas-phase abundance comparable to that in other modeled cores (Figure 12).

Figure 11.

Figure 11. Calculated abundance of important sulfur species in ice, relative to hydrogen.

Standard image High-resolution image
Figure 12.

Figure 12. Calculated abundance of important sulfur species in the gas phase, relative to hydrogen.

Standard image High-resolution image

The abundances of minor sulfur molecules are affected by the proportions of major ice constituents. Molecules whose synthesis is associated with CO and its daughter radicals (primarily atomic C, atomic O is available also from CO2) include OCS, CS, H2CS, and C2S. These are primarily associated with sublayers 1 and 2 (see Figure 13). Oxidized forms SO2 and NS arise in sublayer 3. The above is true for cores CB 17 L, CB 17 S, CB 27, and B68. For CB 26, the diversity is largely limited by the dominating H2S. The efficient synthesis on the surface and subsequent reactive desorption explain why H2S is the most abundant sulfur gas-phase molecule for all modeled cases.

Figure 13.

Figure 13. Calculated abundance, relative to hydrogen, of main sulfur species in ice mantle sublayers for CB 17 S and CB 26. The latter represents a significantly photoprocessed ice.

Standard image High-resolution image

Modeling results show that H2S and SO ices are produced during the freeze-out epoch and contain most of the sulfur reservoir in interstellar ices. The reaction set for sulfur species is rather poor and does not include sulfurous and sulfuric acids (H2SO3 and H2SO4, respectively), which likely are very stable forms of sulfur in watery environments. Oxoacids and their derivatives are possible and non-detectable forms of sulfur in star formation regions (Kalvāns & Shmeld 2010). Given the rich chemistry of sulfur and the radical nature of the SO molecule, it is unlikely that SO is a final, stable, and abundant form of S in interstellar or circumstellar ices. More likely, in terms of this model, it can be viewed as a representation of oxidized sulfur (see also Scappini et al. 2003).

Figure 13 shows an example of depth-dependent composition for sulfur species. H2S and SO are dominant at all depths, except for sublayer 3 at late times, where the abundance of NS is highly enhanced. Although this is probably because of the limited reaction network for sulfur, this illustrates the oxidation of S during longer timescales. Oxidation products are SO, SO2, and NS, or perhaps other species not included in the network. Sulfur in the outer sublayer 1 is less susceptible to oxidation thanks to a higher availability of atomic H.

The high abundance of NS in sublayer 3 for CB 17 L, CB 17 S, CB 27, and B68 is an example of how local conditions in the ice mantle may give rise to peculiar chemical features. NS is produced in all layers via the reactions N + S and NH + S. In sublayers 1 and 2 it is also efficiently destroyed via reactions with atomic C and N. Sublayer 3 is poor with CO and N2. Thus, C and N are unavailable, and NS accumulates after all remaining CO has been converted into CO2.

5. RESULTS: COMPLEX ORGANIC MOLECULES (COMs)

When compared to observational evidence, the synthesis of organic species in the model can be characterized as moderately efficient. For some species, gas-phase abundances are in temporal qualitative agreement with the observations; for others, even ice abundances are well below expectations. To improve this situation, we introduce two simple changes in the model—a modification to CO and H2CO hydrogenation activation energies and, subsequently, a mild and short temperature spike during the quiescent phase. The results of the three simulations are described in the following sections.

To provide an indication of the observed abundances of gas-phase COMs relative to the calculated values, the observed values have been marked in their corresponding figure panels. The observed values have no temporal or spatial constraints with regard to the present model. The 0D character of the simulation means that a qualitative comparison is possible, only. We assume that the synthesis of an organic molecule is reproduced with a degree of success, if its calculated ice abundance is not lower than its observed gas-phase abundance.

5.1. COMs: Standard Model

5.1.1. The Chemistry of COMs

The "Standard" model here is the one described in full in Section 2. Figures 14 and 15 show the calculated abundances in the center of the starless cores, along with known observational values. The species can be categorized in several overlapping groups:

  • 1.  
    those synthesized via the HCO radical, e.g., formic acid HCOOH, formaldehyde H2CO, methyl formate HCOOCH3, formamide NH2CHO, and acetaldehyde CH3CHO;
  • 2.  
    those synthesized via the CH3O radical, e.g., methanol; CH3OH, dimethly ether CH3OCH3, HCOOCH3, and as a subgroup, species whose synthesis involves CH3, which is a daughter of methanol. This subgroup includes CH3CHO, acetonitrile CH3CN, and methane CH4;
  • 3.  
    carbon chain molecules, e.g., CH3CHO, propynone CH2CCO, ketene CH2CO, and acetylene C2H2;
  • 4.  
    species that are typically formed via atom addition on the surface, e.g., CH4, H2CO, HCO, HCOOH CH2CCO, CH2CO, C2H2, methyl imine CH2NH, cyanamide NH2CN, and CH3CN. A relatively high gas-phase abundance during the freeze-out epoch is characteristic for these species.

Figure 14.

Figure 14. Calculated abundances, relative to hydrogen, of observed organic molecules, set No. 1. Gray lines: gas-phase abundance; black lines: abundance in ice (surface and three mantle sublayers). Contraction (Phase 1) ends and the stable core Phase 2 begins at times between 1.4 and 1.5 Myr for all simulations. References: M—Marcelino et al. (2005), O—Öberg et al. (2010), B—Bacmann et al. (2012), C—Cernicharo et al. (2012), V—Vastel et al. (2014).

Standard image High-resolution image
Figure 15.

Figure 15. Calculated abundances, relative to H, of observed and other selected organic molecules, set No. 2. References as in Figure 14.

Standard image High-resolution image

The hydrogenation of CO and that of H2CO are the two most important steps that regulate the general production efficiency for many oxygen-containing COMs in bulk ice. Figures 16 and 17 show the relative abundances of organic species in the four ice layers considered in the models for CB 17 L. This core can be regarded as a "median" case among the five models, without the extremes in ice composition shown by CB 17 S and B68, on one hand, and CB 26 and CB 27, on the other hand (Section 3.1).

Figure 16.

Figure 16. Calculated abundances, relative to H, in ice sublayers for selected organic molecules for the CB 17 L core, set No. 1.

Standard image High-resolution image
Figure 17.

Figure 17. Calculated abundances, relative to H, in ice sublayers for selected organic molecules for the CB 17 L core, set No. 2.

Standard image High-resolution image

Most of the species with abundances shown in Figures 16 and 17 have their peak abundances highest in the outermost sublayer 1. This indicates that even a partial ice sublimation might be sufficient to reproduce the observed gas-phase abundances of many COMs. In sublayer 1, CO is available in vast quantities at any integration time points, while the necessary H atoms are produced in H2O or H2 photodissociation. With the present 0D model it is not possible to determine the percentage of ice mass that should be sublimated. If the modeled ices are assumed to be representative, this proportion would be $\approx 1\%-10$%.

5.1.2. Comparison with Observations

Figures 14 and 15 show that the model predicts relatively high gas-phase abundance for some species in the HCO group—formic acid HCOOH, formaldehyde H2CO, and formamide NH2CHO. Although the observed gas-phase abundances are reached for a short period of time, this occurs during the infall Phase 1 and likely cannot be treated as an agreement with observations of stable starless cores.

For the CH3O group, the calculated abundances are significantly lower than those observed in interstellar conditions. Even assuming instant ice sublimation, observed gas-phase abundances cannot be reproduced for methyl formate HCOOCH3, dimethyl ether CH3OCH3, and methanol CH3OH. This result is similar to that of earlier prestellar core models without (Garrod & Herbst 2006; Garrod et al. 2008) and with bulk-ice photochemistry (Garrod 2013; Paper II). It is this discrepancy that prompts us to proceed with a more detailed investigation on the chemistry of COMs.

5.2. The Compromise Model

5.2.1. Changes Relative to the Standard Model

In the present surface reaction network (Garrod et al. 2008) the activation energy barriers EA for reactions H+CO and H+H2CO are higher than 2000 K, which means that they are inefficient for cold core conditions. Experimental evidence points to much lower values for EA, 300–600 K (Awad et al. 2005; Fuchs et al. 2009). However, putting such low EA values into the present model implies that almost all carbon in ice is in the form of methanol, which is unrealistic. Higher EA is especially required for the CO+H reaction.

We present results of model calculations that use the experimental, low EA value for formaldehyde hydrogenation. For CO hydrogenation EA, an empirical compromise value was used, which lies between the low experimentally detected EA of Fuchs et al. (2009) and the high (2500 K) value proposed by Garrod et al. (2008). Activation energies for the three important CO hydrogenation reactions have been summarized in Table 6. We dub this the "Compromise model" for further reference.

Table 6.  Changes in Reaction Data for the Compromise Model

    Activation Barrier EA, K
No. Reaction Garrod et al. (2008)a Fuchs et al. (2009) Compromise
17 ${\rm H}+\mathrm{CO}\;\longrightarrow \;\mathrm{HCO}$ 2500 390–520 1600
18 ${\rm H}+{{\rm H}}_{2}\mathrm{CO}\;\longrightarrow \;{\mathrm{CH}}_{3}{\rm O}$ 2100 415–490 415
19 ${\rm H}+{{\rm H}}_{2}\mathrm{CO}\;\longrightarrow \;{\mathrm{CH}}_{2}\mathrm{OH}$ 2500 415–490 415

Note.

aThe "standard" values used in this paper.

Download table as:  ASCIITypeset image

5.2.2. Comparison with Observations

Figure 18 shows species whose abundance is substantially affected by the changes in activation energies for reactions 17–19 in Table 6. Comparison with the Standard model results (Figures 14 and 15) reveals that the ice abundance of species that are formed via CH3O is increased by up to six orders of magnitude (CH3OCH3 in B68). Importantly, the ice abundances of CH3OCH3, CH3OH, and CH3O are now above or similar to the detected gas-phase values.

Figure 18.

Figure 18. Calculated abundances, relative to H, of selected organic molecules for the Compromise model. Gray lines: gas-phase abundance; black lines: abundance in ice (surface and three mantle sublayers). For references see the caption of Figure 14.

Standard image High-resolution image

The synthesis of methyl formate HCOOCH3 depends on both HCO and CH3O radicals. We found it impossible to achieve a sufficiently high ice abundance for HCOOCH3 with changes in EA for reactions 17–19. Thus, the calculated relative ice abundance for this molecule remains approximately four orders of magnitude below the observed gas-phase values, which are between 10−11 and 10−9 (Öberg et al. 2010; Bacmann et al. 2012; Cernicharo et al. 2012).

Methyl formate is also inefficiently produced in several other models considering either starless or prestellar cores (Garrod 2013; Vasyunin & Herbst 2013a, 2013b). Chang & Herbst (2014) are able to efficiently synthesize HCOOCH3 with a Monte Carlo model, but their results also show a likely excess of CH4 and a shortage of CO2—discrepancies generally not observed in the above-mentioned other models. These results may indicate that the inefficient synthesis of HCOOCH3 is not an artifact, caused by the limits of the present study. A higher temperature may be necessary for an efficient synthesis of HCOOCH3, as suggested by Garrod & Herbst (2006).

5.2.3. Chemistry Differences with the Standard Model

The most significant abundance increase for the CH3O group is observed in the B68 model. First, the lower temperatures in this core mean a low mobility for atomic H in ice. Second, the abundant CO is now more easily hydrogenated, thanks to the lower EA, and CO soaks up much of the available atomic hydrogen in the mantle. These two factors result in an accumulation of radicals, so that they have a greater chance to react with other multi-atom species. For the present discussion on COMs, the most important radicals are CH3O and HCO. The effect is especially visible for B68 and CB 17 S.

For species that are formed via HCO, the situation is less clear because formaldehyde H2CO is consumed by reactions 18 and 19. In cores CB 17 L and CB 27, formaldehyde has its maximum ice abundance decreased below the detected gas-phase values. A similar situation can be observed for formic acid HCOOH, directly related to HCO.

Figure 19 shows that in the compromise model the abundances of COMs are increased in all sublayers (cf. Figures 16 and 17). The main part of their synthesis still occurs in the outer sublayer 1. Species whose synthesis involves CH3O and CH3 do not reach a steady abundance until the end of the simulation, mainly because they can be regarded as the end products of the CO hydrogenation sequence.

Figure 19.

Figure 19. Calculated abundances, relative to H, in ice sublayers for selected organic molecules, Compromise model, CB 17 L core.

Standard image High-resolution image

5.3. The Synthesis of COMs in Ice in a Temporal Warm-up Event

The discussion in Section 5.2.2 shows that for some COMs (notably, methyl formate) the calculated abundances are nowhere near their observed values. To test how a mild and temporal warm-up event can affect the ice abundances of COMs, we used the B68 Compromise model that undergoes a temperature increase to 20 K for 1 kyr at t = 1.6 Myr.

5.3.1. Chemistry Differences with Previous Models

Figures 20 and 21 show that the temperature spike positively affects the production of a number of solid species—HCOOCH3, CH2CCO, CH2CO, NH2CHO, C2H2, CH4. Notably, the gas-phase abundance is substantially increased for species synthesized on the surface—CH2CCO, H2CO, NH2CN, C2H2, CH3CN, CH2NH, and CH4, thanks to reactive desorption.

Figure 20.

Figure 20. Comparison of calculated COM abundances, relative to H, for the B68 core between the Standard model (st.), Compromise model (c.), and Compromise model with a warm temperature spike (c.w.), set No. 1. The two latter models begin to show differences only after 1.6 Myr, when the temperature spike occurs. References as in Figure 14.

Standard image High-resolution image
Figure 21.

Figure 21. Comparison of calculated COM abundances, relative to H, for the B68 core between the Standard model (st.), Compromise model (c.), and Compromise model with a warm temperature spike (c.w.) at 1.6 Myr, set No. 2. References as in Figure 14.

Standard image High-resolution image

For most species, the changes in ice abundance are within one order of magnitude. However, for methyl formate HCOOCH3 and formamide NH2CHO an increase by approximately five and three orders of magnitude is observed, respectively. We therefore explore more deeply the temperature-spike-induced chemistry of these two species. Initiated at 1.6 Myr, the rise of abundances actually occurs over a period of 100 kyr after the temperature spike. Figure 22 shows the abundances of NH2CHO and HCOOCH3 in ice sublayers.

Figure 22.

Figure 22. Calculated abundances, relative to H, in ice sublayers for species most affected by a 1 kyr 20 K temperature spike in the Compromise model for the B68 core.

Standard image High-resolution image

Methyl formate production occurs mainly in the middle sublayer 2. This is unlike most other COMs in the Standard and Compromise models that are synthesized in the outer sublayer 1 (Figure 19). Sublayer 2 is the only one that is rich in both CO and H2O. The former is the source species for HCO and CH3O radicals that combine into HCOOCH3. Meanwhile, the photodissociation of water is the principal source of atomic hydrogen, necessary for the hydrogenation sequence $\mathrm{CO}\;\longrightarrow \;\mathrm{HCO}\longrightarrow {{\rm H}}_{2}\mathrm{CO}\;\longrightarrow \;{\mathrm{CH}}_{3}{\rm O};{\mathrm{CH}}_{2}\mathrm{OH}$. Thus, CO hydrogenation by the mobile H atoms at the 20 K spike, followed by the combination of HCO and CH3O ≈ 100 kyr after the spike, is the main mechanism for the formation of HCOOCH3.

The case of formamide is different—NH2CHO reaches similar abundance in all three sublayers after the 20 K spike. This is because in none of the sublayers do its main parent species—CO and NH3—both have a high abundance. While NH3 is mainly concentrated in sublayer 3, CO is abundant only in sublayers 1 and 2.

5.3.2. Discussion on Temperature Spike Model Results

The ice abundance increases occur because diffusion, proximity, and activation energy barriers are more effectively overcome at 20 K. The subsequent reactions temporarily reduce ice abundances for the radical species HCO, CH3O, and CH2OH. The results shown in Figures 20 and 21 suggest that even such a small and short warm-up period is sufficient to qualitatively reproduce the abundance of methyl formate in ice and that the ice abundances of other organic species do not become disbalanced. The single-point 0D approach used in the model prevents a quantitative comparison between calculation results and observations.

We suggest two likely causes for a (temporal) temperature increase for a portion of ices in a starless or prestellar core. First, the gas parcel can be exposed to the interstellar radiation because of turbulent motions in the core, as suggested by Boland & de Jong (1982) and Martinell et al. (2006). Second, low-velocity transient shocks within the core (Williams & Hartquist 1984) may also result in mild heating of the gas. Both of these mechanisms also offer a hypothesis for the desorption mechanism—either photodesorption or collisional sputtering of the weakly bound, non-polar ice outer layers. These considerations are supported by the turbulent motions observed in starless cores (Redman et al. 2006; Levshakov et al. 2014; Steinacker et al. 2014).

6. SUMMARY

A model that describes ices in a detailed way was used to investigate ice chemistry in five different examples of starless cores. While the importance of subsurface ice chemistry has been noted in previous studies (Cuppen & Herbst 2007; Kalvāns & Shmeld 2010; Garrod 2013), the most general conclusion from the present study is that chemical processes in ice depend strongly on the particular (sub)layer, where the molecules in consideration are located. This arises because of variations in abundances of major ice components (H2O, CO, CO2, N2, NH3) in different layers on the same grains. This conclusion is in line with the results of a prestellar core model (Paper II) and a pseudo-time-dependent Monte Carlo model by Chang & Herbst (2014). The single sublayer approach (Kalvāns & Shmeld 2010, 2013; Garrod 2013; Belloche et al. 2014) is not suitable for representing subsurface chemistry of interstellar ices. This is especially so if the synthesis of minor species is studied.

The H2O and CO ices in dark cores are photoprocessed to CO2 via sequence (4). This sets limits to core lifetimes, because current observations indicate that the CO2:H2O abundance ratio in ice likely is 44% or less (Boogert et al. 2011). The maximum lifetime for a quiescent and stable molecular core with constant physical conditions (Phase 2) inferred this way in the present model is 650 kyr (CB 17 S core). It is hard to pinpoint a precise value because of the limitations in the physical model and uncertainties regarding the turbulence in dark cores. Taking this into account, we can state that the model indicates dark core lifetimes of <1 Myr. The lower CO2:H2O ratio in Taurus (Whittet et al. 2007) may indicate that these starless cores may be younger than those in most other molecular clouds, observed by Boogert et al. (2011) and Öberg et al. (2011).

The H2O:CO ice abundance ratio cannot be used as a reliable indicator of core age because CO resides mostly on the outer surface of the grains. There, it can be rather easily lost to the gas phase or converted into other species in mild heating events. Such processes can be induced by exposure to interstellar photons or low-velocity shocks within the core, as discussed below.

Because of more efficient desorption, the present model predicts shorter and higher abundance peaks during the cloud collapse phase (or ice formation epoch) for a number of species produced in dense gas or on the surface (Section 3.3). These include H2O, CO2, NH3, H2O2, H2S, SO, SO2, OCS, HCOOH, H2CO, CH2CO, CH4, O2, H2S2, N2, etc.

We find that O2 and H2O2 are the main oxygen reservoirs in ice, except water and carbon oxides. O2 forms in the gas and on the surface and is being converted into the H2O2 via photprocessing in subsurface ice. It was suggested that H2O2 and O2H reach their gas-phase abundance peaks during the freeze-out epoch, thanks to an active molecular exchange between the gas and the surface. The principal cause of desorption, necessary for this exchange, is irradiation by interstellar photons.

Calculations show that the HNC:HCN ratio in ice can be as high as 1.75 (in the CB 27 core model), while it remains close to or below unity in the gas phase. Sulfur in ice is roughly equally divided between reduced (H2S) and oxidized (SO) forms. Subsurface processing of these sulfur species gives rise to OCS, NS, and H2CS, indicating an interesting diversity for the sulfur chemistry.

Investigating the chemistry of COMs, we found that lower activation energies—more consistent with experimentally detected values—for hydrogenation reactions of CO and H2CO may help explain the observations of COMs in the interstellar medium. A mild and temporal warm-up event helps to produce an abundance of methyl formate in ice that is higher and more consistent with gas-phase HCOOCH3 observations. Such an event also substantially increases the ice abundance of formamide.

Desorption by interstellar photons and ice sputtering in grain collisions may transport the COMs to the gas phase. Both of these mechanisms require either turbulence within the cores or external influence. These findings are probably supported by the fact that two of the dense cores with detected COMs—L1689b (Bacmann et al. 2012) and B1-b (Marcelino et al. 2005; Öberg et al. 2010; Cernicharo et al. 2012)—have nearby star formation regions (Kirk et al. 2007; Öberg et al. 2010), while the third, L1544 (Vastel et al. 2014), is known to be turbulent and on the verge of collapse (Caselli et al. 2002; Tatematsu et al. 2014).

I acknowledge the support of the Ventspils City Council. This research has made use of NASA's Astrophysics Data System.

Please wait… references are loading.
10.1088/0004-637X/806/2/196