1 Introduction

Around 200 different molecules have been detected in interstellar space, not counting isotopologues, as listed on the Cologne Database for Molecular Spectroscopy.Footnote 1 One of the motivations for studying the formation of organic molecules in interstellar medium is to trace potential chemical pathways to the formation of biologically important molecules (Garrod and Widicus Weaver 2013). Models based on the gas-phase chemistry alone are unable to reproduce abundances of many complex molecules observed in cold molecular clouds and their cores (van Dishoeck 2014). In particular, such species as methanol as well as many more complex species are effectively formed on the surface of dust grains (Watanabe and Kouchi 2002). The gas-phase abundance of these molecules can be enhanced by orders of magnitude due to ejection of these species from icy mantles of dust grains in shocks (Lefloch et al. 2017).

Shocks are produced in the interstellar medium in a variety of sources. (i) During a supernova event the matter thrown away forms a shock wave that propagates in the interstellar medium and interacts with molecular clouds. Studies of molecular emission from the clouds encountered by a supernova remnant can provide information about magnetic field strength, density and velocity of the gas (Wardle and Yusef-Zadeh 2002). (ii) Complex organic molecules (COMs) have been detected in outflows powered by low-mass and high-mass protostars—it is suggested that these species are ejected during the sputtering of grain mantles in shocks (Palau et al. 2017). (iii) It has been proposed by Requena-Torres et al. (2006) that the sputtering of ice mantles of dust grains in shocks may be responsible for the ‘rich’ chemistry observed in the Galactic centre clouds.

The interesting question is the degree of processing of gas material and grain mantles in the shock, the survival of molecules in the hot postshock gas (Holdship et al. 2017; Palau et al. 2017). The result can be a considerably different chemical composition in the shocked gas than observed in quiescent clouds (Bergin et al. 1998). In order to relate the emission line spectra to the properties of the shocked gas such as gas density and abundances of chemical species, it is necessary to treat comprehensively the dynamic evolution of the gas, the chemistry of the medium, and physical processes of ion and molecule excitation and de-excitation (Flower and Pineau des Forêts 2015).

Different types of shocks can be distinguished depending on the value of the magnetosonic speed in the interstellar gas (Flower 2007; Draine 2011). If the shock speed is higher than any signal speed in the shocked medium, the jump (J-type) shock forms. In J-type shocks, physical conditions at shock front change in a discontinuous way, leading to dissipation of the flow kinetic energy in a thin region. As a consequence, high peak temperatures and full dissociation of molecules take place. At low shock speeds, magnetosonic waves precede the shock, and the coupling between ions and neutral gas results in a continuous change in physical parameters of the gas. The continuous (C-type) shocks are formed in this case. Given that H2 molecule is the main coolant of the medium, C-type shocks can exist only up to a limiting value of the shock speed at which the collisional dissociation of H2 (and other molecules) takes place in the hot postshock gas. For typical physical conditions inside dark molecular clouds, the transition from C-type to J-type shock takes place at shock speeds about \(40\mbox{--}60~\mbox{km}\,\mbox{s}^{-1}\) (Draine and McKee 1993; Le Bourlot et al. 2002). Models of magnetohydrodynamic non-dissociative shocks in dense molecular clouds have been created by a number of authors, e.g. Mullan (1971), Draine (1980), Draine et al. (1983), Kaufman and Neufeld (1996), Wardle (1998), Guillet et al. (2007), van Loo et al. (2009), Flower and Pineau des Forêts (2015). Numerical models of shock waves usually consider in detail either the gas dynamics, but reduced chemical network is used, or vice versa—the parametric model of the steady state profile of the shock is used to study in detail chemical evolution of the gas (Holdship et al. 2017). The advantage of a magnetohydrodynamic model over the parametric one is that a large number of physical parameters (e.g. cosmic ray ionization rate, intensity of interstellar background radiation field, dust properties, and etc.) may be varied. Here, we present a magnetohydrodynamic model of C-type shock coupled to a full gas–grain chemical network.

The calculations consist of two steps. In the first step, we compute the chemical composition of a dense cloud with a fixed density. The abundances of chemical species, gas and dust temperatures obtained during this step are used as the initial chemical and physical state of the gas for the shock simulations.

2 Parameters of the dark cloud

In this section, physical parameters of the molecular cloud are described. We use the set of ‘low-metal’ initial abundances of elements except for He, C, and N (Graedel et al. 1982). For He we use a value of 0.09 with respect to hydrogen (Wakelam and Herbst 2008). For C and N the abundances close to that observed in \(\zeta\) Oph diffuse cloud are used (Jenkins 2009; Hincelin et al. 2011). The C/O elemental ratio in non-refractory material in dense clouds is not well known, and here we adopt the C/O ratio of 0.5 (Hincelin et al. 2011). The species are assumed to be initially in atomic form except for hydrogen, which is assumed to be molecular. The initial elemental fractional abundances relative to the total H nucleus number density are given in the Table 1, along with the choice of ionization state. All species are assumed to be in the gas phase.

Table 1 Initial elemental abundances with respect to H nucleus number density

The measured values for the cosmic ray ionization rate in dense interstellar gas lie in the wide range from \(10^{-17}~\mbox{s}^{-1}\) to values as high as \(10^{-15}~\mbox{s}^{-1}\) (Dalgarno 2006). A scatter in measured values may be due to details of the measurements, the physical and chemical models used in the analysis of observational data, and may also reflect intrinsic variations of the cosmic ray flux from cloud to cloud. Magnetic field effects can significantly reduce the cosmic ray ionisation in dense cloud cores (Padovani and Galli 2011).

The dispersion of turbulent velocities (micro-turbulence speed) and gas velocity gradient are necessary parameters for radiative transfer calculations. The micro-turbulence speed determines the excess in line width over the thermal value while the gas velocity gradient determines the length of the region where molecular radiation is coupled to the ambient gas. Starless cores in molecular clouds—sites of low-mass star formation—present spectra of core-tracing species that have close-to-thermal line widths (Tafalla et al. 2004; André et al. 2014). Here, the dispersion of turbulent velocities is taken equal to the sound speed in the gas at 10 K—\(0.2~\mbox{km}\,\mbox{s}^{-1}\). Molecular clouds are not rigorously characterized by large-scale systematic motion, as required for the large velocity gradient approximation to be valid. However, the characteristic value of the velocity gradient can be estimated as the ratio of cloud line width (in velocity units) and cloud radius. From this, one can estimate the velocity gradient of the order of \(1~\mbox{km}\,\mbox{s}^{-1}\,\mbox{pc}^{-1}\) (Goldsmith 2001).

The nature and structure of shock waves travelling through molecular clouds are strongly dependent upon the strength of magnetic field \(B_{0}\) (Draine 1980). According to the analysis of the data on magnetic field strength in molecular clouds by Crutcher (1999), an approximate empirical relation holds between line-of-sight magnetic field strength and gas density \(B_{\mathrm{{los}}} = \beta n_{\mathrm{ {H,tot}}}^{1/2}\), where \(n_{\mathrm{{H, tot}}}\) is the total hydrogen nuclei number density, \(\beta\approx0.7~\upmu\mbox{G}\,\mbox{cm}^{3/2}\). We assume that \(B_{0}/B_{\mathrm{{los}}} \simeq 1.5\mbox{--}2\).

A ‘classical’ single-size grain model is considered. The grains, made of silicate material, are assumed to be spherical particles with a radius of \(0.1~\upmu\mbox{m}\) and internal density of \(3.5~\mbox{g}\,\mbox{cm}^{-3}\). The dust–gas mass ratio is taken equal to 0.01. The grains are initially bare. We neglect the change in the grain radius by freeze out of gas-phase molecules onto dust grains. Optionally, our computer code is able to treat polycyclic aromatic hydrocarbon (PAH) molecules, but no PAH molecules are considered in our ‘standard’ model. Optical properties for silicate and carbonaceous spheres are used from (Draine and Lee 1984; Laor and Draine 1993; Li and Draine 2001; Weingartner and Draine 2001a) and are available at website of Prof. B.T. Draine.Footnote 2

The initial ortho-/para-H2 ratio is taken to be 1 as a representative value of dark molecular clouds at young ages, \(t < 1~\mbox{Myr}\) (Pagani et al. 2013). The chemistries of ortho- and para-H2 are not distinguished in our model.

The physical parameters of our ‘standard’ model are given in the Table 2. The description of the simulations of gas-phase and grain surface chemistries, calculations of level populations of ions and molecules, heating and cooling processes, and shock structure are given in Appendices A, B, C, and D, respectively. The list of chemical reactions of collisional dissociation of species is given in Appendix E.

Table 2 Parameters of the dark cloud

3 Results

3.1 Chemical evolution of the dark cloud

The objectives of the modelling of chemical evolution of a static dark cloud are: (i) verification of our chemical model; (ii) evaluation of the chemical composition of the gas before the shock wave propagates through it. In simulations of chemical evolution of a static dark cloud, a simple zero-dimensional model is considered in which density remains fixed as the chemistry progresses from initial specimen abundances. In this section the main results of our calculations are presented, and the comparison is made with the results of other workers.

3.1.1 General results

The temperature of the neutral gas component is shown on the Fig. 1a. During the evolution of the dark cloud all gas components have close temperatures. The gas reaches a thermal equilibrium after about \(10^{3}~\mbox{yr}\) and gas temperature remains at relatively constant level of about 13–15 K. The main heating mechanism of the gas is the cosmic ray driven chemistry, \(\mathcal{G}_{\mathrm{{chem}}}\), see Fig. 1b. The main cooling mechanisms are cooling via molecular and atomic line emission, \(\mathcal{G}_{\mathrm{{n,rad}}}\), and by gas–dust collisions, \(\mathcal{G}_{\mathrm{{n, d}}}\). At evolutionary times \(t > 10^{5}~\mbox{yr}\), heavy species become adsorbed on dust grains (see Fig. 1c). But total gas-phase depletion of heavy species does not take place due to reactive desorption mechanism. At gas densities \(n_{\mathrm{{H,tot}}} \gtrsim 10^{4}~\mbox{cm}^{-3}\), the gas and dust start to couple thermally via collisions and molecular depletion effect on the gas temperature diminishes, see also Goldsmith (2001).

Fig. 1
figure 1

Chemical and thermal evolution of the static cloud: (a) gas temperature; (b) rates of main heating and cooling mechanisms of the gas, the term \(\mathcal{G}_{\mathrm{{n,rad}}}\) includes cooling by ions CI, CII, OI and molecules H2, CO and H2O; (c) abundances of gas species relative to hydrogen nuclei; (d) abundances of grain mantle species, s-X denotes species X in icy mantles, \(\mbox{C}_{n}\mbox{H}_{m}\) denotes all hydrocarbon molecules with \(n \geq2\)

The main ice mantle constituent of interstellar grains is H2O ice (Fig. 1d). According to our simulations, water on the grain surface is mainly formed in the reaction between H atom and hydroxyl OH, with some contribution of other channels. In our model, the contribution of CO2 to icy mantles is approximately 10–20 per cent that of the H2O ice at \(10^{5}\mbox{--}10^{6}~\mbox{yr}\). Carbon dioxide forms on the grain surface via reactions CO + OH and H + HOCO (HOCO in turn is also produced in reaction CO + OH). The dominance of one or the other reaction channel in CO2 formation strongly depends on the adopted parameters such as thickness and height of activation barriers of chemical reactions. The abundance of methanol is relatively low, 0.1–1 per cent that of water abundance at \(10^{5}\mbox{--}10^{6}~\mbox{yr}\). Hydrogen atoms adsorbed on dust grains participate in numerous surface reactions. As a result, hydrogen molecule formation via direct association of hydrogen atoms is negligible except at very late evolutionary times \(t > 3 \times10^{7}~\mbox{yr}\). According to our simulations, hydrogen molecule is produced through hydrogen abstraction reactions on grain surface from molecules HCO, H2CO, HNO, and other species, see also Tielens and Hagen (1982), Hasegawa et al. (1992). The abundances of species in ice mantles of grains found in our simulations are in reasonable agreement with astronomical ice observations (Boogert et al. 2015).

The temperature of dust grains is equal to 9.3 K according to our calculations. At \(A_{\mathrm{{V}}} \gtrsim3\), dust grains are mainly heated by interstellar radiation field at infrared wavelengths. Most of dust grains are either neutral or have the charge −1. According to our simulations, the photoelectric emission by cosmic ray induced UV radiation and collisional attachment of ions and electrons are all important in grain charging as ionization fraction has reached values of \(10^{-7}\mbox{--}10^{-8}\). Analogous results were found by Ivlev et al. (2015). The effect of cosmic ray induced UV radiation field on grain charge decreases with increasing gas density (Guillet et al. 2007).

3.1.2 Comparison with NAUTILUS model

The comparison is made between the results of our calculations and the NAUTILUS code (Ruaud et al. 2016), see Fig. 2. In NAUTILUS simulations, we set identical to our model the initial elemental abundances (Table 1), parameters of grain surface chemistry (Table 3), gas temperature evolution with time (Fig. 1a). The difference between the results of our simulations and that of NAUTILUS code is no higher than an order of magnitude for most simple species. Note that different data on binding energies and different databases for gas-phase chemical reactions are used in our calculations and in NAUTILUS code.

Fig. 2
figure 2

Comparison of our simulation results with NAUTILUS results. The results of simulations are also shown with updated chemical network according to Chabot et al. (2013) and using the data on specimen binding energies as in NAUTILUS model. The results of NAUTILUS model with direct cosmic ray desorption reduced by an order of magnitude are shown

Table 3 Parameters of grain surface chemistry

The method of direct cosmic ray desorption of adsorbed species by Hasegawa and Herbst (1993) is incorporated in NAUTILUS code, whilst the method by Roberts et al. (2007) is used in our model, see Appendix A. The cosmic ray desorption of volatile molecules such as CO is very efficient according to NAUTILUS simulations, and the abundance of CO in the gas phase stays at high level. If the cosmic ray desorption is reduced by an order of magnitude, the results become much closer, see Fig. 2. According to our results, the decrease of the ionization fraction of the gas starts at early evolutionary times compared with NAUTILUS results. At about \(10^{2}\mbox{--}10^{3}~\mbox{yr}\), large unsaturated carbon-chain molecules Cn and CnH and their anions are produced in significant amounts in our model, and the gas is effectively neutralized through reactions:

$$ \mbox{C}_{n}^{-} + \mbox{C}^{+} \to \mbox{C}_{n} + \mbox{C}, \qquad \mbox{C}_{n} \mbox{H}^{-} + \mbox{C}^{+} \to \mbox{C}_{n} \mbox{H} + \mbox{C}. $$
(1)

Chabot et al. (2013) provided a set of branching ratios for the reactions involving carbon-chain species. For the reaction \(\mbox{C}_{n}^{-}+\mbox{C}^{+}\), internal energy of intermediate complexes is high, and three fragment channels are dominating. These branching ratios are so far not taken into account in the UDfA chemical network but are included in the latest version of the KIDA network (Wakelam et al. 2015) that is used in NAUTILUS. New branching ratios prevent formation of large carbon-chain molecules at early times (Chabot et al. 2013). Indeed, our results on electron abundance become very close to NAUTILUS results when we update our chemical network according to data by Chabot et al. (2013), see Fig. 2.

Vasyunin et al. (2004) analysed the influence of errors in the rate constants of gas-phase chemical reactions on the calculated specimen abundances. They found that errors in the abundances of simple species lie within 0.5–1 order of magnitude. Wakelam et al. (2006) found similar uncertainties and discussed the differences between two wide used gas-phase chemical networks. Recently, Penteado et al. (2017) presented a systematic study of the effect of uncertainties in the binding energies on abundances of chemical species. They found that there is a large variation in the abundances of ice species when binding energies are varied within their errors. Our results on CH3OH abundance in icy mantles become close to those by NAUTILUS model, if the same data on binding energies are used in both models, see Fig. 2.

3.2 Shock model results

According to the Herschel Gould Belt survey studies of nearby star-forming clouds, the typical lifetime of starless cloud cores with density \(\sim10^{4}\mbox{--}10^{5}~\mbox{cm}^{-3}\) is about \(10^{6}~\mbox{yr}\) on average (André et al. 2014). Moreover, best agreement between observed and modelled specimen abundances in cloud cores is most often achieved at ‘early’ times of \(10^{4}\mbox{--}10^{6}~\mbox{yr}\) (Wakelam et al. 2006; Penteado et al. 2017). For the shock wave modelling, the gas chemical composition and gas temperature at 0.5 Myr are chosen.

3.2.1 Comparison with the shock model by Flower and Pineau des Forêts (2015)

A comparison is made between the results of our simulations and the results obtained using the code mhd_vode by Flower and Pineau des Forêts (2015). The similar specimen abundances are used at the start of both simulations. Figure 3 shows velocities and temperatures of gas components, and gas cooling rates due to emission in molecular and atomic lines. The general behaviour of physical parameters is similar in our model and that by Flower and Pineau des Forêts (2015)—both models provide approximately the same shock width, maximal temperatures of gas components, and the same maximal velocity difference between ion and neutral fluids. The shock model by Flower and Pineau des Forêts (2015) does not consider grain surface chemistry and includes simpler gas-phase chemical network than our model—the influence of expanded chemistry on shock structure is minimal.

Fig. 3
figure 3

Comparison between results of our calculations, lower panels, and the C-type shock structure calculated by the model by Flower and Pineau des Forêts (2015), upper panels, for a shock velocity \(u_{\mathrm{{s}}}= 20~\mbox{km}\,\mbox{s}^{-1}\). The integration of differential equations starts at the zero point of the \(z\)-axis. The rates of thermal energy transfer in collisions from the gas to molecules and ions are shown (‘collisional’), the net rate of radiative energy loss is also given for H2 molecule (‘radiative’)

The rates of collisional energy transfer from the gas to the molecules and ions are shown in the Fig. 3, the net rate of radiative energy loss is also given for H2 molecule (assuming that radiative transitions are optically thin). The main gas coolant in hot shocked gas is H2 molecule. As the shocked gas cools, the contribution of other molecules to the cooling process becomes significant. The gas cooling by molecules OH, NH3 and CH3OH is not yet taken into account in our model. Cooling by these species may be significant in the case of elevated abundances (Flower et al. 2010). In the postshock region, rotational levels of ground vibrational state of H2 are overpopulated due to slow de-excitation rates of these levels. In this case, the collisional energy transfer from the gas to the H2 molecule changes the sign, and collisions of gas species with H2 heat the gas, see Fig. 3. Our model produces low rate of radiative energy loss by H2 in the cold gas where H2 molecules can be vibrationally excited only by formation processes—this effect is not taken into account in our model.

3.2.2 Chemical evolution of the shocked gas

Figure 4 shows shock structure and evolution of abundances of simple molecules that are often considered as shock tracers. There is an increase of abundances of icy mantle species at the beginning of the shock before sputtering starts—the grain velocity decreases and their number density increases in the magnetic precursor as the grains are (partially) coupled to the ion fluid. The sputtering of grain mantles starts when the gas–grain relative speed reaches about \(5~\mbox{km}\,\mbox{s}^{-1}\). Thus, the threshold shock speed for grain mantle sputtering is about \(10~\mbox{km}\,\mbox{s}^{-1}\). The main sputtering projectiles are heavy species He and CO at low shock speeds. These results are similar to those by Jiménez-Serra et al. (2008), Van Loo et al. (2013). Molecules released from grain mantles are chemically processed in the hot postshock gas. As the gas cools, high abundances of species produced in the shock persist in the postshock gas until the time-scale for the individual molecule to deplete onto dust grains. The simulation results on abundances of simple species are in agreement with the findings by other workers, see e.g. Bergin et al. (1998), Charnley and Kaufman (2000), Viti et al. (2011), Flower and Pineau des Forêts (2012).

Fig. 4
figure 4

Shock structure and evolution of abundances of simple species. Results are shown for four shock velocities: \(10, 20, 30, 45~\mbox{km}\,\mbox{s}^{-1}\). Each row of graphs corresponds to the shock velocity shown on the corner of the first graph in the row

Figure 5 shows the evolution of abundances of some COMs in the shock wave. Species that are produced in ‘nonenergetic’ atom addition reactions (e.g. \(\mbox{CH}_{3}\mbox{OH}\), \(\mbox{CH}_{3}\mbox{OCH}_{3}\)) are abundant in icy mantles in the preshock gas. Abundances of many other complex species (e.g. \(\mbox{HCOOCH}_{3}\), \(\mbox{C}_{2}\mbox{H}_{5}\mbox{OH}\)) are low, as radical–radical association reactions that produce such species are inefficient at low dust temperatures. In shock, the increase of ion–neutral drift velocity is rapid, and gas reaches the maximum temperature soon after the sputtering. Hence, the sputtering of grain mantles takes place in the region close to the temperature peaks of neutral gas and ions. At high shock speeds, molecules are destroyed in the hot shocked gas via reactions with H atoms and collisional dissociation reactions. The survival time of complex molecules in the hot shocked gas is low—of the order of dozens of years, see Fig. 5.

Fig. 5
figure 5

Evolution of abundances of some COMs in the shock: methanol \(\mbox{CH}_{3}\mbox{OH}\), formic acid HCOOH, isocyanic acid HNCO, acetaldehyde \(\mbox{CH}_{3}\mbox{CHO}\), dimethyl ether \(\mbox{CH}_{3}\mbox{OCH}_{3}\). Results are shown for shock velocities \(20, 30 \mbox{ and } 45~\mbox{km}\,\mbox{s}^{-1}\)

At high shock speeds, there is high abundance of atomic hydrogen in the gas produced in collisional dissociation reactions. It triggers formation of hydrocarbon molecules on the grain surface. At shock speed \(45~\mbox{km}\,\mbox{s}^{-1}\), the abundance of methanol in icy mantles of dust grains in the cool postshock gas equals to 10–15 per cent relative to water ice, while in the preshock gas it constitutes only about 0.5 per cent, see Fig. 5. According to observational data, the abundance of methanol ice (relative to H2O ice) ranges from upper limits of no more than a few percent toward dense molecular clouds to substantial levels of up to 30 per cent toward a few young stellar objects, e.g. RAFGL7009S and W 33A (Dartois et al. 1999; Pontoppidan et al. 2003). The methanol production in the cooling postshock gas is one of the possible explanations of high abundance variations of methanol ice observed in astronomical sources. Methanol is reformed in the gas phase via production on the grain surface followed by the desorption into the gas and, with a minor contribution, via gas-phase reactions. The efficiency of reactive desorption mechanism is a key parameter that controls methanol re-formation in the gas phase in the cooling postshock region. At shock speed \(45~\mbox{km}\,\mbox{s}^{-1}\), the peak abundance of gas-phase methanol in the postshock gas is \(8 \times10^{-8}\) at standard model parameters, and is about \(7 \times10^{-9}\) at efficiency of reactive desorption of \(f = 0.001\)—an order of magnitude lower than our standard value.

COMs are effectively produced in the gas phase in the postshock region. At about 104 years after the passage of high speed shock (\(45~\mbox{km}\,\mbox{s}^{-1}\)), the abundance relative to H nuclei of methyl formate HCOOCH3 (both in the gas and in icy mantles) is about \(4 \times10^{-10}\) that is almost three orders of magnitude higher than in the preshock gas. The analogous effect is seen for ethanol \(\mbox{C}_{2}\mbox{H}_{5}\mbox{OH}\), but the abundance of ethanol reaches low values of about 10−11 in our model. The abundance of acetaldehyde \(\mbox{CH}_{3}\mbox{CHO}\) in icy grain mantles is low as it reacts rapidly with hydrogen atoms. In postshock region, acetaldehyde is effectively produced in the gas phase, it abundance reaches values of about \(10^{-9}\)—an order of magnitude higher than in the preshock gas. The main parent species in the gas-phase synthesis of COMs are \(\mbox{H}_{2}\mbox{CO}\), \(\mbox{CH}_{3}\mbox{OH}\), \(\mbox{C}_{2}\mbox{H}_{4}\) and radicals \(\mbox{CH}_{3}\), \(\mbox{CH}_{3}\mbox{O}\), \(\mbox{C}_{2}\mbox{H}_{5}\).

4 Discussion

The reactions of ‘nonenergetic’ atom addition on the grain surface play a main role in dark cloud chemistry as adsorbed species are not able to cross large reaction barriers (Charnley and Rodgers 2008; Fedoseev et al. 2017). Parameters that have a strong influence on the simulated grain mantle composition are the adopted binding and diffusion energies of species and activation energies of grain surface reactions (Taquet et al. 2012; Penteado et al. 2017). The chemical composition of icy mantles in dark clouds strongly depends on the grain temperature, cosmic ray ionization rate, gas density and evolution stages experienced by the gas prior to significant grain mantle formation. The problem of chemical evolution of a particular dark cloud and shock propagation through it must be considered jointly in order to make reasonable predictions on specimen abundances and molecular line intensities.

Most grain surface reactions are exoergic. Part of the energy released in a reaction will be immediately transferred to the ice mantle of dust grain, other part is transformed to the kinetic energy and internal (electronic and ro-vibrational) excitation of reaction products (Lamberts et al. 2014; Fredon et al. 2017). We have not included dust heating by chemical reactions. Dust heating by this mechanism is negligibly small at dark cloud conditions. However, an explosive release of the chemical energy stored in icy mantles as free radicals may take place as the dust temperature rises in the shock—this mechanism was considered by Shen et al. (2004) in the study of cosmic ray induced explosive chemical desorption in dense clouds. Quite possible that this mechanism can lead to the liberation of volatile species from icy mantles to the gas phase before sputtering in the shock. At the model parameters in question, the effect of dust heating on the shock chemistry is small—the dust temperature is increased up to 15 K in the shocked gas before grain mantle sputtering, but the gas passes quickly through this region. The possible influence of dust heating on the grain surface chemistry in the shock is beyond the scope of the present paper.

Observations of molecular outflows of protostars indicate that molecules show two different kinds of profiles, with CH3OH, NH3 and other species emitting only at relatively low outflow velocities, whereas H2O shows bright emission even at the highest velocities (Codella et al. 2010; Gómez-Ruiz et al. 2016; Holdship et al. 2016). It is explained by chemical modelling—molecules that have destruction reactions with low activation energy are destroyed in hot gas (Viti et al. 2011). The discussion on the behaviour of simple molecules in C-type shocks and their potential to be a shock tracer was given by Holdship et al. (2017). According to our simulations, the time-scale for molecule survival in the high-speed shock is low—of the order of dozens of years. Recently, Palau et al. (2017) studied the evolution of COMs using the parametric shock model by Jiménez-Serra et al. (2008). They found that COMs can survive long enough after the passage of a shock. Our results imply that more strict constraints must be put on the physical parameters for the shock regions where COMs are observed—shock velocity and gas density must be low enough to allow COMs survive in the hot shocked gas. Other possibility is that observed COM’s emission comes from a postshock region where these molecules have been reformed in the gas (Codella et al. 2015; Palau et al. 2017). It is likely that a fraction of molecules could be destroyed in the sputtering process, but we do not consider this effect in our model, see discussion by Suutarinen et al. (2014).

The COMs destroyed in the hot shocked gas could be reformed in the postshock region. The high abundance of simple species and radicals that are produced in the sputtering of icy mantles and collisional dissociation reactions is one of the factors that promote COM production in the postshock gas. The abundance of H atoms is other important factor controlling the chemistry in the interstellar gas (Cuppen et al. 2009; Wakelam et al. 2010). As the gas cools, H atoms drop out the gas through the adsorption on dust grains. It is shown that the high abundance of H atoms in the cooling postshock gas may trigger formation of methanol on the grain surface. Indeed, as was shown by Cuppen et al. (2009), the increase of H/CO ratio in the gas in dark molecular clouds shifts the grain surface chemistry to the formation of complex hydrocarbon molecules. In the postshock region, gas-phase methanol is mainly produced via reactive desorption mechanism. The observations of methanol in the cold gas that has experienced the shock passage in the recent history can be used to quantify the magnitude and importance of reactive desorption mechanism. The methanol is the parent specimen in the production of other COMs and the gas-phase abundance of such species strongly depends on methanol abundance.

One of the shortcomings of our model is the simple dust model. The size distribution of grains has significant effect on the chemical evolution of the dark cloud (Iqbal and Wakelam 2018) and on the shock structure and the gas-phase chemistry in the shock (van Loo et al. 2009; Flower and Pineau des Forêts 2012). Grain shattering in shocked gas may produce numerous small grain fragments, which increase the total dust grain surface area. The effect is that C-type shocks become shorter and warmer, which in turn affects the chemistry and molecular emission (Anderl et al. 2013). Grain shattering and its feedback onto the dynamics of C-type shocks are found to be significant at densities higher than about \(10^{5}~\mbox{cm}^{-3}\) (Guillet et al. 2011).

The chemical network used in our model is incomplete for proper modelling of chemistry of COMs and extensions of gas-phase and grain surface chemistries must be used in future work (Garrod 2013; Taquet et al. 2016; Fedoseev et al. 2017).

5 Conclusions

The evolution of abundances of COMs in C-type shock wave is considered. The main results of the paper are summarized below:

  1. (i)

    The sputtering of grain mantles takes place in the shock region close to the peak of neutral gas temperature. As a result, time-scale for molecule survival in the high-velocity shock is low—of the order of dozens of years.

  2. (ii)

    For high-speed shocks, \(u_{\mathrm{{s}}} \gtrsim40\mbox{--}45~\mbox{km}\,\mbox{s}^{-1}\) at preshock gas density \(n_{\mathrm{{H,tot}}} = 2 \times10^{4}~\mbox{cm}^{-3}\), the abundance of H atoms and radicals in the postshock gas is relatively high, that affects the gas-phase and grain surface chemistries. The efficient methanol production on the surface of dust grains in the cool postshock gas may be one of the reasons of high abundance of methanol ice observed toward some young stellar objects.

  3. (iii)

    Gas-phase methanol is re-formed in the postshock gas via reactive desorption mechanism. The efficiency of reactive desorption is a key parameter that determines the gas-phase abundance of methanol and other complex species that are produced via it.

  4. (iv)

    At high shock speeds, the postshock abundances of some COMs such as methyl formate may be much higher than in the preshock gas due to efficient gas-phase production.

A comparison has been made between our simulations and results obtained with other publicly available codes: NAUTILUS for modelling chemistry of dark clouds (Ruaud et al. 2016) and mhd_vode for modelling shocks (Flower and Pineau des Forêts 2015). The difference between the results of our simulations and those of NAUTILUS code is not larger than an order of magnitude for most simple species. There is a good agreement between our results and that of the mhd_vode code—both codes predict similar profiles of gas temperature and velocity, and similar shock widths.

The shock model presented can be employed to interpret observational data on molecular emission from outflows powered by protostars, to study formation of biologically relevant molecules in shock regions, and can be also used for modelling of cosmic masers. The influence of various physical parameters on the chemistry and molecular emission of the shocked gas can be studied using this model.