Niclas A.
West‡
a,
Lok Hin Desmond
Li
a,
Tom J.
Millar
b,
Marie
Van de Sande
c,
Edward
Rutter
a,
Mark A.
Blitz
a,
Julia H.
Lehman§
a,
Leen
Decin
d and
Dwayne E.
Heard
*a
aSchool of Chemistry, University of Leeds, Leeds, LS2 9JT, UK. E-mail: d.e.heard@leeds.ac.uk
bAstrophysics Research Centre, School of Mathematics and Physics, Queen's University Belfast, University Road, Belfast BT7 1NN, UK
cSchool of Physics and Astronomy, University of Leeds, Leeds, LS2 9JT, UK
dInstituut voor Sterrenkunde, KU Leuven, Celestijnenlaan 200D, 3001 Leuven, Belgium
First published on 24th February 2023
Rate coefficients for the reaction of CN with CH2O were measured for the first time below room temperature in the range 32–103 K using a pulsed Laval nozzle apparatus together with the Pulsed Laser Photolysis–Laser-Induced Fluorescence technique. The rate coefficients exhibited a strong negative temperature dependence, reaching (4.62 ± 0.84) × 10−11 cm3 molecule−1 s−1 at 32 K, and no pressure dependence was observed at 70 K. The potential energy surface (PES) of the CN + CH2O reaction was calculated at the CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ level of theory, with the lowest energy channel to reaction characterized by the formation of a weakly-bound van der Waals complex, bound by 13.3 kJ mol−1, prior to two transition states with energies of −0.62 and 3.97 kJ mol−1, leading to the products HCN + HCO or HNC + HCO, respectively. For the formation of formyl cyanide, HCOCN, a large activation barrier of 32.9 kJ mol−1 was calculated. Reaction rate theory calculations were performed with the MESMER (Master Equation Solver for Multi Energy well Reactions) package on this PES to calculate rate coefficients. While this ab initio description provided good agreement with the low-temperature rate coefficients, it was not capable of describing the high-temperature experimental rate coefficients from the literature. However, increasing the energies and imaginary frequencies of both transition states allowed MESMER simulations of the rate coefficients to be in good agreement with data spanning 32–769 K. The mechanism for the reaction is the formation of a weakly-bound complex followed by quantum mechanical tunnelling through the small barrier to form HCN + HCO products. MESMER calculations showed that channel generating HNC is not important. MESMER simulated the rate coefficients from 4–1000 K which were used to recommend best-fit modified Arrhenius expressions for use in astrochemical modelling. The UMIST Rate12 (UDfa) model yielded no significant changes in the abundances of HCN, HNC, and HCO for a variety of environments upon inclusion of rate coefficients reported here. The main implication from this study is that the title reaction is not a primary formation route to the interstellar molecule formyl cyanide, HCOCN, as currently implemented in the KIDA astrochemical model.
There is a desire to understand what aspects of the reaction potential energy landscape control the behavior of the reaction rate coefficient and the branching to products as a function of temperature. As discussed in a recent perspective,3 both PESs with small barriers to reaction and/or submerged barriers to reaction can cause sharp increases in rate coefficients at low temperatures until in some cases the collision limit is reached. For the case of small barriers to reaction under low temperature reaction conditions, weakly bound complexes in the entrance channel for reaction are formed with less energy compared with higher temperatures. This causes their lifetime for dissociation back to reactants to become sufficiently long that quantum mechanical tunneling through reaction barriers to products can become competitive, leading to a dramatic increase in the rate coefficient with decrease in temperature.3 Two examples are the reaction of OH with Complex Organic Molecules (COMs) such as CH3OH,3,5–8 and the reaction of C(3P) with H2O.9 For the case of submerged barriers, an increase in the rate coefficient with a decrease in temperature is observed consistent with a negative activation energy to reaction. Two examples are the reaction of C2H with O2,10 and the reaction of O(1D) with CH4.11 If a calculated PES is used in conjunction with kinetic models to predict k(T), any uncertainty in the height or width of relatively small potential energy barriers can lead to large uncertainties in the prediction of reaction rate coefficients. Therefore, these reaction-specific PES attributes are important to closely examine alongside robust experimental data spanning a broad temperature range, especially including low temperatures.
The current study focuses on the low temperature reaction of the cyano radical with formaldehyde, CN + CH2O (Reaction 1), which was suggested by astrochemists to be the primary formation route to the interstellar molecule formyl cyanide, HCOCN,12 a suggestion reinforced recently by the ab initio quantum calculations of the PES and master equation calculations of the rate coefficients by Tonolo et al.13 Chemical intuition and results from past studies of similar reactions14–16 suggest, however, that CN will likely attack CH2O via a hydrogen abstraction mechanism, forming hydrogen cyanide (HCN) and the formyl radical (HCO). Both reactants and predicted hydrogen abstraction products have already been observed in several astrochemical environments, such as Asymptotic Giant Branch (AGB) stellar winds,17,18 the InterStellar Medium (ISM),19–26 and dark clouds,27–30 providing further motivation for the current study at low temperatures. Additionally, the CN + CH2O reaction is relevant to combustion processes,31 planetary atmospheres,32 and Titan's atmosphere.33,34 The inclusion of new kinetic data for neutral–neutral reactions at very low temperatures can make a significant difference for the abundance of key species in interstellar environments.3,35
Previous experimental work on the CN + CH2O reaction measured overall rate coefficients between 297–673 K by Yu et al.36 and 294–769 K by Chang and Wang37 (for ease of reading these two works are referred to from now on as YCW36,37) using the Pulsed Laser Photolysis–Laser-Induced Fluorescence (PLP–LIF) technique in resistively-heated flow cells. In these studies, k1(T) was found to decrease slightly with decreasing temperature, from approximately 4 × 10−11 cm3 molecule−1 s−1 near 770 K to approximately 1.5 × 10−11 cm3 molecule−1 s−1 at room temperature. The uncertainties quoted for these studies were on average 2.8% for Yu et al.,36 and 1.2% for Chang and Wang.37 It was predicted that a hydrogen abstraction process, forming HCN, was likely to dominate the chemical mechanism. Indeed, a prior theoretical study (QCISD/6-31G**//UHF/6-31G**) found only a very small (∼2.7 kJ mol−1) barrier to this H atom abstraction reaction.38 In comparison, a very recent calculation by Tonolo et al.13 (CCSD(T)/CBS+CV) reports a small submerged barrier for the abstraction pathway forming HCN. Further aspects of the reaction potential energy surface were explored in this calculation, with particular attention paid to addition reaction pathways forming CN–CH2O adducts, with bonds formed between the C of formaldehyde and either end of CN. However, the high-pressure limit rate coefficient (2 × 10−10 cm3 molecule−1 s−1) calculated using transition state theory and based on this new PES13 is close to an order of magnitude larger than the previously measured room temperature studies of YCW. Exothermic product channels for the reaction of CN with CH2O include:13,39,40
![]() | (1a) |
![]() | (1b) |
![]() | (1c) |
![]() | (1d) |
![]() | (1e) |
Enthalpies for reactions (1a)–(1c) are taken from Ruscic and Bross 2019;39 for reaction (1d) from Born et al.;40 but for reaction (1e) the zero-point energy corrected energy change calculated by Tonolo et al.13 is given. In the work presented here, low-temperature reaction rate coefficients were measured for the first time below 294 K for the CN + CH2O reaction using the PLP–LIF technique in a pulsed Laval nozzle apparatus. Using CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ, an ab initio PES was calculated for CN + CH2O and utilized in the Open Source MESMER software package41 to calculate temperature-dependent rate coefficients and product branching ratios. Using a fitting procedure, the parameters within MESMER were optimized to give the best agreement with experimental data over the range 32–769 K. These MESMER rate coefficients were then used to develop a parameterisation via a modified Arrhenius (MA) equation and extrapolated to 4 K. The fits of the rate coefficients were then incorporated into astrochemical simulations of cold dense clouds, AGB stars, and hot molecular cores.
The generation of uniform, thermalized, cold, gas flows for kinetics measurements with this apparatus was described previously in detail and will only be described briefly here.14,42–47 Gaseous formaldehyde (CH2O) is a difficult reagent to work with and particular care was given to its preparation from paraformaldehyde. Preparation of mixtures with the bath gas, its handling, and quantification of its concentration are necessary for accurate kinetics studies. Generation of the CH2O reagent gas was performed by controllably heating paraformaldehyde powder (Sigma-Aldrich, 95%) in an evacuated 500 mL glass bottle (Duran) with a heat gun (Steinel, model HL1810S) to ∼70 °C. During heating of paraformaldehyde, the nascent CH2O gas was then passed through a −10 °C cold trap. The cold trap was submerged in ethanol (VWR, 99.96%) which was chilled with a refrigerated immersion probe (LaPlant, model 100CD). The cold trap was utilized to trap any water or other condensable byproducts generated during the heating of paraformaldehyde. The purified CH2O was then allowed to flow into evacuated cylinders until reaching a pressure of ∼200 torr (∼26.7 kPa). The cylinders were then filled with argon (BOC, 99.998%) or nitrogen (BOC, 99.998%) to ∼6 atm (∼608 kPa) total (generating ∼4.4% mixtures of CH2O/bath gas). The gas in the cylinders was then allowed to mix for >12 hours. This method of generating CH2O is similar to methods utilized previously by our group and other groups.14,48,49 To generate the cyano radical, CN, the vapor from the solid precursor cyanogen iodide (ICN, Acros Organics, 98%, vapor pressure ∼1 torr (∼133 Pa) at 298 K.50,51) was entrained in ∼2 atm (∼203 kPa) of Ar or N2 bath gas.
The controlled mixing of reagent gases was accomplished through the use of Mass Flow Controllers (MFC) (MKS, type 1179A) such that final mixtures of gases were ∼0.1–1% CH2O, ∼0.005% ICN, and ∼99% Ar or N2 bath gas with a total flow rate of ∼2–5 slm. The gases were then allowed to mix in a mixing manifold prior to being pulsed through the Laval nozzle. Since CH2O was found to slowly repolymerize, depositing solid paraformaldehyde in the gas lines and causing the MFC calibration to slowly drift, the concentration of CH2O was determined via UV absorption measurements by sampling gas from the tubing between the mixing manifold and the pulsed valves for each concentration of CH2O utilized in a kinetics experiment.14 Measurements of the concentration of CH2O were made using a custom-made 1 m path length UV/Vis absorption cell filled to ∼1.2 atm (∼122 kPa) with the gas-mixture as measured by a capacitance manometer (MKS, 0–100 PSIA (∼689 kPa)). The light source for absorption measurements was a UVB lamp (EXOTERRA, UVB200) with continuous output around ∼290–350 nm. Absorption measurements were performed with a UV/Vis spectrometer (Ocean Optics, HR4000CG-UV-NIR) with 0.75 nm resolution. Absorption spectra were integrated for ∼2 seconds and 4 spectral traces were averaged. Representative averaged UV absorption spectra for CH2O are shown in Fig. S1 in the ESI.† Using concentrations determined from this spectrometer, a typical half-life of the generated ∼4.4% CH2O in an N2 bath in the cylinders was determined to be approximately 4 days.
After the mixing manifold, the gas mixtures were sent to two solenoid valves (Parker, Series 9) to be pulsed at 5 Hz into the pre-expansion region upstream of the Laval nozzle. Gas in the pre-expansion reservoir underwent a controlled expansion through a custom-made, axisymmetric, converging-diverging Laval nozzle such that the flow into the vacuum chamber at 0.3–2 torr (40–267 Pa) was a wall-less reactor. The vacuum chamber was continuously evacuated by two Roots blowers operating in parallel: a Roots blower (Leybold RUVAC 251) backed by a rotary pump (Leybold D65B) and a Roots blower (Edwards EH250) backed by a rotary pump (Edwards ED660) and the pressure in the vacuum chamber was monitored by a capacitance manometer (Leybold, type CTR90, 0–10 torr (0–1.3 kPa)). Flow temperatures between 32–103 K were achieved by switching between Laval nozzles of Mach numbers between 2.49 and 5.00 and/or switching between Ar and N2 bath gases. Impact pressure measurements using a Pitot tube were utilized to determine the Mach number along the flow, and via the Rayleigh equations, the density and rotational–translational temperature of the flow,4 as illustrated for temperature in Fig. S2 (ESI†).
In order to initiate reactions of CN + CH2O, the ICN precursor was photolyzed at 266 nm (∼30 mJ per pulse) using the fourth harmonic of a Nd:YAG laser (Quantel Q-Smart 850), which was directed co-linearly with the supersonic flow along the axis of the Laval nozzle (“Pump laser” in Fig. 1), generating a uniform density of CN radicals. Although the CN radical is initially generated with rotational excitation,52–54 collisions with the bath gas at the densities used ensure that rotational relaxation occurs on a timescale that is short compared with the timescale for removal of CN due to its reaction with CH2O. The time evolution of the CN radicals produced are monitored using laser-induced fluorescence following excitation of the R1(2) rotational line of the B2Σ − X2Σ (1,0) vibronic transition at 357.893 nm, which is generated by a Nd:YAG (Quantel, Q-smart 850) pumped dye laser (Sirah, Cobra Stretch, with Pyridine 2 dye). This “probe” laser beam (Fig. 1) crossed the gas flow perpendicularly at a fixed position ∼10–25 cm from the nozzle exit depending on the stable flow length of the particular nozzle used. Fluorescence from CN (B2Σ − X2Σ) was focused through a series of lenses and through a bandpass filter at 400 nm with a FWHM of 40 nm (Thorlabs, FB400-40) onto a Channel PhotoMultiplier (CPM) (PerkinElmer, C1952P). The gain of the CPM was controlled with a custom-built time-dependent high voltage gating module in order to block scattered light from the pump laser at 266 nm. Digitization and integration of the CPM signal were performed on an oscilloscope (LeCroy, Waverunner LT264), which is then transferred and saved to a computer for further analysis via a LabVIEW program. The timing of the experiment was also controlled via LabVIEW communication with a digital delay generator (BNC, Model 555), which randomly varied the order of the time delays between the pump and probe laser after each gas pulse.
From the generated PES, statistical rate theory calculations were performed using the MESMER software program41 in order to obtain the rate coefficients of the system. The stationary points of the ab initio calculations provide the energies, rotational constants, and vibrational constants required by the MESMER input file. The energy wells along the PES are divided into energy grains, where each grain couples the reactant, intermediate, and product species to one another via the microcanonical rate coefficients, k(E). The individual grains can be populated/depopulated by exchange with other grains via collisional energy transfer with the buffer gas. The microcanonical rate coefficients were calculated with either Rice, Ramsperger, Kassel, and Marcus (RRKM) theory65 for reactions involving a defined transition state or the inverse Laplace transformation (ILT) method66 for barrierless reactions. Collisional energy transfer probabilities were described using the exponential-down model.67 Corrections for quantum mechanical tunneling were also included using the Eckart expression.68 The set of coupled differential equations that describe each of the energy grains is known as the energy grained master equation (EGME) and can be described by:
![]() | (2) |
p = UeΛtU−1p(0) | (3) |
The initial photolytic production of CN and fast collisional relaxation of rotationally-excited52–54 CN into the CN (X2Σ, N = 2) laser-probed state led to a rapid rise in the CN LIF signal (≲1 μs) which was not resolved in our experiments. Therefore, traces were analyzed after ∼5 μs following the initial rise in the signal, after which there was an exponential decay of CN due to diffusion, reaction with CH2O, and reaction with other possible species. Diffusion of CN out of the volume of supersonic flow through which the pump laser passed is given by:
![]() | (4) |
The reaction of CN with CH2O (reaction 1, bimolecular rate coefficient k1), and potential reactions with other species (for example, with the precursor, other photolysis products, or reagent impurities) is given by:
![]() | (5) |
![]() | (6) |
![]() | (7) |
![]() | (8) |
![]() | ||
Fig. 2 Averages of ≥5 temporal decays of the CN integrated LIF signal utilized to determine the pseudo-first-order rate coefficients, kobs, for loss of CN at 32 K, both in the absence of CH2O and for [CH2O] = 6.06 × 1013 molecule cm−3, at a total density of (3.24 ± 0.24) × 1016 molecule cm−3 in Ar bath gas, together with exponential fits (eqn (7)) to the data. No data could be recorded for longer pump-probe time delays due to the limited reaction time within the low-temperature uniform supersonic flow prior to the breakup of the flow. |
For each CH2O concentration, the experiment was repeated at least five times and the kobs values obtained were averaged to give obs. Second-order (bimolecular) plots of
obsversus [CH2O] were then generated at each temperature, examples of which are given in Fig. 3 (with the intercept kint subtracted for clarity) with a linear least-squares fit of eqn (6) used to determine k1 from the gradient.
![]() | ||
Fig. 3 Variation of the intercept-subtracted, average loss rate of CN, ![]() ![]() |
For the linear fits of the second-order plots such as shown in Fig. 3, care was taken to include only the range of [CH2O] for which obs is linear with [CH2O] to avoid any possible influence from CH2O dimers. Evidence for dimerization of CH2O was found in previous low-temperature studies at higher [CH2O] performed using a Laval nozzle, namely for rate coefficient determinations of the reactions OH + CH2O,15 CH + CH2O,14 and NH2 + CH2O.70 The range of values of the intercept (at [CH2O] = 0 molecule cm−3) was kint = 2200–7600 s−1, and the variation of
obs with [CH2O] without the subtraction of kint can be found in Fig. S3 of the ESI.† We have been very cautious in the maximum [CH2O] used to obtain kobs for a given temperature, always using less than the [CH2O] at which second-order plots have been seen to become non-linear in similar studies of radical + CH2O reactions.14,15 The values of k1(T) determined in this work for the CN + CH2O reaction, together with corresponding experimental conditions, are given in Table 1. At 70 K, k1 was determined at two overall densities providing evidence that k1 is independent of pressure.
T (K) | Bath gas | N total (1016 molecule cm−3) | k 1(T)b (10−11 molecule−1 cm3 s−1) |
---|---|---|---|
a Pitot tube measurements of impact pressures were utilized to determine T and Ntotal. Here 1σ fluctuations of these values in the cold flow along the axis of the nozzle are reported.
b The error of each k1(T) value represents the standard error in the fitted value of the gradient of ![]() |
|||
32 ± 2 | Ar | 3.24 ± 0.24 | 3.57 ± 0.53 |
32 ± 2 | Ar | 3.24 ± 0.24 | 4.62 ± 0.84 |
40 ± 4 | Ar | 8.36 ± 1.19 | 2.26 ± 2.11 |
53 ± 4 | Ar | 7.04 ± 0.74 | 3.18 ± 1.19 |
56 ± 6 | Ar | 7.58 ± 1.11 | 1.56 ± 1.94 |
70 ± 11 | Ar | 11.18 ± 2.54 | 1.51 ± 0.49 |
70 ± 2 | N2 | 2.91 ± 0.20 | 1.30 ± 0.52 |
84 ± 3 | N2 | 7.56 ± 0.63 | 0.99 ± 0.27 |
92 ± 6 | N2 | 4.99 ± 0.74 | 1.80 ± 0.29 |
103 ± 10 | N2 | 6.80 ± 1.57 | 1.45 ± 0.19 |
The measurement uncertainty for k1 for CN + CH2O given in Table 1 is larger than reported in our determination of the rate coefficient for the CH + CH2O reaction.14 The uncertainty quoted is a statistical error only from linear fits of the type shown in Fig. 3, and adding systematic errors did not increase the overall error significantly. The primary reason is that k1(T) is about 10–100 times smaller than kCH+CH2O(T) across the range of temperatures studied (0.99 × 10−11–4.62 × 10−11versus 4.86 × 10−10 – 11.15 × 10−10 cm3 molecule−1 s−1, respectively). Hence, for the same [CH2O] and available reaction time (controlled by the length of the stable flow for a given Laval nozzle), the LIF signal from CN does not decay as much as for CH, resulting in a relatively larger uncertainty in the fit of eqn (7) to the data to determine kobs. The values of k1(T) determined in this work for T = 32–103 K for the CN + CH2O reaction are shown in Fig. 4, and also in Fig. S8(a) and S10 of the ESI,† together with experimental measurements from two other groups36,37 reported over the temperature range 294–769 K.
![]() | ||
Fig. 4 Measured values of k1(T) versus temperature from this work (black squares), Chang and Wang37 (blue triangles), and Yu et al.36 (green triangles) (collectively YCW). The error bar of each k1(T) from this work are those given in Table 1, and the YCW values are assigned an error equal to 0.064 × k1(T), see text for details. The Laval fit (black line) are the MESMER simulations which are a best-fit to the experimental Laval measurements only, with ab initio parameters (Table 2) including a submerged transition state. For details of the Laval + Lit 1 and Laval + Lit 2 fitting scenarios and parameters see Table 2. The inset shows the extrapolation of Laval (black line) and Laval + Lit 2 (red line) models to lower temperatures, where significant deviation in these models is only observed below 30 K. See text and Table 2 for further details of these scenarios. |
Fig. 4 shows that k1(T) exhibits a weak positive temperature dependence above room temperature but a strong negative temperature dependence below ∼100 K, with a likely minimum somewhere between 100–200 K, suggestive of a change in reaction mechanism between the low and high temperature regimes. The values of k1(T) at low temperatures are less precise than those previously reported at higher temperatures,36,37 due to the challenges for this experiment as discussed above. It is noted though that for the previous data reported at higher temperatures, the absolute concentration of CH2O was not determined by UV absorption spectroscopy (in contrast to the low temperature work reported here), rather manometric/flow methods were used to determine [CH2O].
The geometries of the stationary points, obtained at the M06-2X/aug-cc-pVTZ level of theory, are shown in Fig. S4 (ESI†). The optimized Cartesian coordinates, vibrational frequencies, energy values, and additional scans along the reaction potential energy surface can be found in the ESI.†
A weakly bound van der Waals complex VDW (−13.3 kJ mol−1) is identified following the approach of CN to CH2O. As seen in Fig. 5, this complex leads to four different product pathways, namely H atom abstraction to form HCN (P1), H atom abstraction to form HNC (P2), and addition of CN onto the O atom (P3–4) eventually forming either NCO + 3CH2 (P3) or HC(O)CN + H (P4). The formation of HCN + HCO (P1, −163.4 kJ mol−1) is accessible through the submerged barrier TS_VDW/P1 (−0.62 kJ mol−1) while the formation of HNC + HCO (P2, −103.6 kJ mol−1) involves a small positive energy barrier relative to the CN + CH2O entrance channel (TS_VDW/P2, 3.97 kJ mol−1). The error of these two calculated barrier heights, even with the high level of theory used here, are such that they could both be either positive or submerged barriers if calculated at a different level of theory. At the CCSD(T)/aug-cc-pVTZ level of theory, the error of the energy is estimated to be 3.0–4.5 kJ mol−1 (250–450 cm−1).1
The addition of CN onto the O atom of CH2O involves surmounting a large barrier TS_VDW/P3-4 (32.9 kJ mol−1), leading to the formation of the intermediate H2C–O–CN (Int1, −119.2 kJ mol−1). H2C–O–CN can dissociate to form NCO + 3CH2 (P3, 174.4 kJ mol−1) or undergo cyclization through TS_1/2 (−44.2 kJ mol−1) to form the cyclic intermediate Int2 (−81.2 kJ mol−1). By going through TS_2/3 (−62.5 kJ mol−1), the ring opens to form the intermediate H2C(O)CN (Int3, −153.4 kJ mol−1). Breaking one of the CH bonds gives the products HC(O)CN + H (P4, −66.7 kJ mol−1). Although the IRC calculation did not converge successfully for TS_3/P4 (−41.0 kJ mol−1), judging from the vibrational mode of the imaginary frequency, it is likely that it is the TS connecting Int3 and P4. A dashed line connecting TS_3/P4 reflects the incomplete mapping of the IRC along this coordinate.
Relaxed scans at the BHandHLYP/aug-cc-pVDZ and M06-2X/aug-cc-pVTZ level of theory attempted to map out the approach of the two reactants, CN and CH2O. Two different reactant approaches were investigated: the CN radical approaching from the oxygen side of CH2O, and CN approaching from the hydrogen side. For each scan, the distance between the two reactants was fixed while all other coordinates were allowed to optimize. When CN approaches from the oxygen side of CH2O, as shown in Fig. S5 (ESI†), it is favorable for CN to orient the carbon towards CH2O at the beginning of the approach due to the dipole–dipole attraction. The potential energy decreases smoothly as the distance between CN carbon and CH2O oxygen decreases, eventually reaching the potential energy well where the van der Waals structure VDW is located.
In the case where CN approaches from the hydrogen side of CH2O, as shown in Fig. S6 (ESI†), it is favorable at first for CN to orient itself such that the nitrogen side points towards CH2O. The potential energy decreases with decreasing separation between the two moieties until encountering a fairly “flat” region of the potential energy surface. Here, with a distance of approximately 3.3 Å between nitrogen of CN and carbon of CH2O, or a center of mass distance between the moieties of approximately 4.5 Å apart, the overall energy of the system (BHandHLYP/aug-cc-pVTZ, uncorrected for ZPVE) reaches −5 kJ mol−1 relative to the entrance channel. In order to further explore this “flat” region of the PES, a relaxed scan from this point has been done using the OC⋯N angle as the scanning parameter, as shown in Fig. 6.
While the CN radical rotates around CH2O, the potential energy first experiences a fairly flat region of the potential and then falls smoothly into the potential energy well corresponding to the van der Waals structure VDW. Thus, it is suggested that both ways of approach can eventually lead to the van der Waals structure VDW.
A direct H abstraction mechanism of CN from CH2O to form HCN + HCO was studied in a previous calculation using the QCISD/6-31G**//UHF/6-31G** level of theory.38 This prior work suggested that the reaction mechanism involved overcoming a small positive ∼2 kJ mol−1 (emerged) barrier relative to the CN + CH2O entrance channel. The current study and a very recent study13 have not identified this pathway on the reaction coordinate, perhaps because of the lower level of theory used for geometry optimization (UHF/6-31G**) in the previous study.13 In the current study, instead of direct abstraction, an indirect channel to form HCN + HCO was found which involves the van der Waals complex VDW and a small submerged barrier TS_VDW/P1. A very recent theoretical study13 by Tonolo et al. also identified the VDW structure on the reaction PES (−13.0 kJ mol−1 relative to reactants, including ZPE, CCSD(T)/CBS + CV) leading to hydrogen abstraction to form HCN + HCO through a submerged barrier (−1.26 kJ mol−1). This pathway is consistent with our current work, with less than 0.5 kJ mol−1 difference in energy in the complex and barrier. Our work also details for the first time the presence of the HNC pathway from the VDW complex.
The work from Tonolo et al. identifies several stationary points along the reaction coordinate not used in the current study. For example, Tonolo et al. claim a barrierless route to forming a C–C bond directly from the reactants, resulting in a tetrahedral intermediate which they label 1C in a deep potential well (−153 kJ mol−1). Although we agree that this structure can be formed (Int3 in Fig. 5), we were unable to connect this structure to the reactants (relaxed scans, BHandHLYP/aug-cc-pVDZ and M06-2X/aug-cc-pVTZ) without surmounting a large barrier at least 50 kJ mol−1 (CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ) higher in energy than the reactants, as shown in Fig. S7 (ESI†). It is likely that the overestimated rate coefficients from Tonolo et al. compared to previous and current experimental work are directly related to not identifying a substantial barrier to C–C bond formation in forming their intermediate 1C. This structure 1C and the remainder of the addition pathways Tonolo et al. identified coming from this structure, including roaming mechanism pathways, were not considered in our search for reactions which would occur at temperatures relevant to the current study. This impacts on the relative importance stated for the pathway leading to the formation of formyl cyanide (HCOCN), although we identified a new pathway to HC(O)CN + H through VDW which does not involve roaming and has a barrier (TS_VDW/P3-4).
In comparison with the PES of OH + CH2O, which has been studied extensively in previous work,2,16,71–75 the results from the current work on CN + CH2O show that both systems share a similar shape of the PES in terms of the energy profile or mechanism for the H abstraction reaction. Following the approach of the two reacting species, a pre-reaction complex is formed followed by a transition state with a small barrier to form products. The difference in energy which determines whether the barrier is positive or submerged relative to the reactants energies is in the kJ mol−1 range, and hence within the uncertainty of most calculation methods, and so whether the transition state is submerged or not will depend on the level of theory used.
For example, the energy values of the transition state for H abstraction channel of OH + CH2O computed with more robust methods tend to give lower values,2 decreasing the barrier from being slightly emerged to slightly submerged. The latest value reported from Machado et al.2 obtained at CCSD(T)/CBS level is approximately −5.7 kJ mol−1.
As CN + CH2O did not show a pressure dependence for k1(T), either experimentally near 1017 molecule cm−3 or from MESMER simulations between 1015–1018 molecule cm−3, the gas density was set at 1013 molecule cm−3 (so making sure in a pressure independent region) for the MESMER simulations. Above 1019 molecule cm−3 and T < 50 K, a pressure dependence was evident and the VDW species was populated. No pressure dependence was observed in the current calculations, implying that the van der Waals complex is not significantly stabilised, even at the lowest temperatures.
Therefore, to a very good approximation, Fig. 5 can be reduced to just the formation of HCN and HNC as shown in Fig. 7 without losing chemical information; reducing the system to just HCN formation would still describe the system to better than 99% of the product yield.
![]() | ||
Fig. 7 A simplified PES used for MESMER fitting scenarios, see Table 2, together with molecular structures at key stationary points. Initial van der Waals complex, VDW red, formation is followed by transition states to products, HCN + HCO (P1, blue) and HNC + HCO (P2, green) respectively. See text for details. |
The inverse Laplace transformation (ILT) method66 was used to calculate the microcannonical rate coefficients for the barrierless formation of the VDW complex from CN + CH2O. The ILT approach overcomes the problem of explicitly assigning a transition state for a barrierless process, whose position would be varying as a function of temperature. The ILT method is especially convenient when experimental data are available, as for this study. The ILT parameters for VDW formation were assigned by:
![]() | (9) |
As well as performing simulations, MESMER can adjust the important rate controlling parameters in eqn (9), as well as the transition state energies in order to best fit to available experimental data. During such data fitting, MESMER uses the Marquardt algorithm76 to adjust the parameters in order to minimize χ2:
![]() | (10) |
Name of fitting scenario | Laval | Laval + Lit 1 | Laval + Lit 2 |
---|---|---|---|
A ∞ILT,VDW and n∞ILT,VDW are the ILT parameters defined in eqn (9) and the other parameters are the energy of the transition state (TS_VDV/P1, see Fig. 5 and 7) and its imaginary frequency.a Units are cm3 molecule−1 s−1.b Units are cm−1.c N is the number of degrees of freedom, the number of data points minus the number of fitting parameters. The reported errors are 1 sigma. | |||
Temperature range (K) | 32–103 | 32–769 | 32–769 |
A
∞ILT,VDW ×10−11![]() |
5.29 ± 0.8 | 2.5 ± 0.8 | 6.58 ± 0.7 |
n ∞ILT,VDW (unitless) | −0.77 ± 0.17 | 0.016 ± 0.005 | −0.10 ± 0.02 |
TS_VDW/P1 (kJ mol−1) | −0.62, fixed | −0.62, fixed | 4.0 ± 0.9 |
Imaginary frequencyb (cm−1) | 215, fixed | 215, fixed | 806 ± 24 |
χ 2/Nc | 1.09 | 4.24 | 0.80 |
Initially, MESMER data fitting was carried out with energies fixed at the ab initio calculated values, as shown in Fig. 7, with only the ILT parameters in eqn (9) being adjusted. When fitting to just the low temperature Laval nozzle data (Laval scenario in Table 2) an excellent visual fit with the data was obtained (see Fig. 4). χ2/N is a measure of the goodness of fit, where χ2 is given by eqn (10) and N is the number of degrees of freedom, which is equal to the number of data points minus the number of fitting parameters. A χ2/N value close to 1.0 represents a good fit, and from Table 2 it can be seen that the Laval model is close to 1.0, and indicates the experimental errors in Table 1 are realistic.
When MESMER simultaneously fits to the low and high-temperature data of YCW36,37 (denoted Laval + Lit 1 scenario in Table 2), the fit to the data is much worse, with χ2/N = 4.24, and as can clearly be seen in Fig. 4. Therefore it is concluded that MESMER calculations based on our ab initio calculated values is not capable of reproducing all the experimental data shown in Fig. 4. The reason is because the ab initio value for the energy of the transition state (TS_VDW/P1) is submerged, i.e. is negative with respect to the reagent energies, and a reaction mechanism proceeding via a submerged transition state is going to predict k1 to decrease with increasing temperature, which is at odds with the high temperature literature data, which requires the transition state to have a positive energy with respect to the reagents. The simplest way to account for the variation of k1 across the full range of temperatures is to increase the transition state energy so that it is positive, making it consistent with the high-temperature literature data trends and to increase the imaginary frequency of the tunnelling coordinate in the transition state so that quantum mechanical tunnelling overcomes the effect of the positive barrier, leading to an increase in k1 at lower temperatures, as is observed experimentally. When the model scenario Laval + Lit 2 (Table 2), where both the energy and the imaginary frequency of both transition states for the channels producing HCN and HNC products (TS_VDW/P1 and TS_VDW/P2, respectively in Fig. 7) are adjusted in the MESMER simulations using a best-fit procedure, the k1 experimental data across the full range of temperatures can be fitted well, evidenced by a value of χ2/N = 0.80. The optimum MESMER adjustments from the fitting were an increase in the energy from −0.62 to + 4.0 kJ mol−1 and an increase in the imaginary frequency from 215 to 806 cm−1 for TS_VDW/P1. We stress that the 4.59 kJ mol−1ab initio energy difference between the two transition states leading to HCN and HNC products (TS_VDW/P1 and TS_VDW/P2 in Fig. 7, respectively) is maintained, with the energy of the transition state for HNC (TS_VDW/P2) now being +8.59 kJ mol−1, and the imaginary frequency for the transition-state leading to HNC was also increased to ∼806 cm−1. This is reasonable given the similar calculated geometries of the two transition states, but with just the C and N flipped, as shown in Fig. S4 (ESI†), and the tunnelling motion is therefore roughly the same.
The MESMER simulation model scenario Laval + Lit 2 of the fit to the data is shown in Fig. 4, and is much improved over Laval + Lit 1. In order for the Laval + Lit 2 scenario to give the best fit to the k1 data, the transition state energy needed to be increased by ∼4.6 kJ mol−1 (from −0.62 kJ mol−1 to +4.0 kJ mol−1 for TS_VDW/P1, which is still the TS via which the reaction occurs to form HCN products) from our ab initio calculated values at the CCSD(T)/aug-cc-pVTZ level of theory.1 However, this adjustment seems reasonable given it is close to the estimated uncertainty at the CCSD(T)/aug-cc-pVTZ level of theory of ∼3.0–4.5 kJ mol−1.1 Similarly, even though the imaginary frequencies has been increased significantly, from 205 cm−1 to 806 cm−1, the adjusted value of ∼800 cm−1 again is not unreasonable.5 As both transition states emanate from the same pre-reaction complex leading to H abstraction products, it seems reasonable to adjust the energies of each of these by the same amount whilst maintaining the same difference between them. The motivation for changing the initial ab initio results to the fitted results using MESMER is to show that the experimental data can be fitted well if the potential energy surface, specifically for these two transition states, is modified. Just changing the properties of the two initially formed transition states which are formed from the same pre-reaction complex is the most straightforward way to do this.
From Fig. 4 it can be seen that the minimum in the rate coefficient k1 occurs around 150 K, which is the point at which the controlling influences of quantum mechanical tunnelling and the 4.0 kJ mol−1 barrier for TS_VDW/P1 become balanced. Also included in Fig. 4 is the simulation of model Laval down to 7 K, where at 10 K the predicted value of k1 is approximately twice that of the Laval + Lit 2 model. This significant difference emphasizes how sensitive k1 is to the parameters used in the models when extrapolating down to very low temperatures, i.e. <10 K, as both the Laval and Laval + Lit 2 models give almost the same quality fits to the experimentally measured k1 using the Laval nozzle.
The equation utilized commonly as a parameterisation in several astrochemical models (which use, for example, the UMIST Rate12 (UDfa, https://udfa.ajmarkwick.net) or KIDA (https://kida.obs.u-bordeaux1.fr78) databases) to represent k(T) is a single modified Arrhenius (MA) equation:
![]() | (11) |
In the parameterization given by eqn (11) each MESMER data point was assigned a 5% error. The values for the best-fit parameters from eqn (11) fits to the Laval + Lit 2 simulation data are given in Table 3. From Table 3, it is noted that χ2/N is always less than 1, which implies that on average this representation is better or within 5% of the MESMER simulation data, see Fig. S10 (ESI†). Also included in Table 3 are the parameterization values for the Laval model for the range T = 3–80 K. But it is the parameterization of the Laval + Lit 2 model that represents our recommendation of k1(T) for use in astrochemical modelling.
T range (K) | α (cm3 molecule−1 s−1) | β (Unitless) | γ (K) | χ 2/Nb | MESMER fitting scenario |
---|---|---|---|---|---|
a We recommend using the Laval + Lit 2 fitting scenario for astrochemical modelling. b N is the number of degrees of freedom, which is equal to the number of MESMER simulated data points minus the number of fitting parameters. | |||||
3–80a | (3.05 ± 0.14) × 10−12 | −1.11 ± 0.02 | −2.18 ± 0.23 | 0.70 | Laval |
4–20 | (1.18 ± 0.08) × 10−11 | −0.58 ± 0.03 | 0.53 ± 0.24 | 0.55 | Laval + Lit 2 |
20–100 | (3.72 ± 0.15) × 10−12 | −1.09 ± 0.04 | 5.2 ± 1.8 | 0.12 | Laval + Lit 2 |
100–300 | (5.99 ± 0.14) × 10−11 | −2.19 ± 0.04 | −313.4 ± 7.7 | 0.11 | Laval + Lit 2 |
300–1000 | (6.26 ± 0.30) × 10−11 | −0.02 ± 0.03 | 398 ± 16 | 0.03 | Laval + Lit 2 |
As a check of whether the parameterisation for k1(T) gives values that are realistic at the very lowest temperatures, classical capture theory (CCT) calculations have been carried out for the CN + CH2O reaction to calculate the rate coefficient at the collision limit kcoll(T), using the molecular parameters given in Table S11 (ESI†). As shown in Fig. S9 (ESI†) for the CN and CH2O collision pair, kcoll(T) is dominated by large dipole–dipole interactions below 600 K. As shown in Fig. 8, kcoll at ∼3 K is equal to 1.3 × 10−9 cm3 molecule−1 s−1, and from the MESMER simulations k1(T) is still calculated to be a factor of 10 or so less than kcoll(T), and does not get close nor indeed exceed kcoll(T) for all temperatures relevant to conditions used in astrochemical models.
![]() | ||
Fig. 8 Plot of the classical capture theory (CCT) rate coefficient (blue curve) for the CN + CH2O reaction versus the k1(T) from the Laval + Lit 2 model (black filled circles), see Table S10 (ESI†), where errors have been propagated via the covariance matrix obtained from fitting to the experimental data, and the MA extrapolation to the lowest temperature (red curve). Note that the predicted values of k1(T) increase markedly at low T, but never approach the values from CCT. |
The inclusion of the title reaction does, however, impact on the abundance of HCOCN which our ab initio calculations rule out as a product of the CN + CH2O reaction. Previous modeling13 that included this product channel with a rate coefficient of up to 2 × 10−10 cm3 molecule−1 s−1 had shown that the calculated HCOCN fractional abundance was similar to that observed, 3.5 × 10−11, in the cold dark cloud TMC-1, (n(H2) = 104 cm−3, T = 10 K).80 Our results indicate that, since CH2O is an abundant reactant, the total production rate of HCOCN is likely to be an order of magnitude less than the observational requirement.80
We find that changing the parameterisation of the reaction rate to the new measured and calculated rates does not affect the chemistry of both C-rich and O-rich outflows. The abundances and column densities of all species involved (CH2O, CH, HCO, HCN and HNC) are not changed. Fig. S13 and S14 (ESI†) show the abundances of HCO and HCN calculated for the high mass-loss rate O-rich and C-rich outflow, respectively. This behaviour is also seen in the lower mass-loss rate outflows.
Using the MESMER package the product branching ratio was calculated and the formation of HNC was not found to be important as a reaction channel. MESMER was then used to simulate the rate coefficients k1(T) from 4–1000 K, and this dataset was used to recommend a parameterization from best-fit modified Arrhenius expressions for k1(T) for use in astrochemical modelling. Even at 4 K the simulated values of k1(T) were a factor of 10 less than the collision limit. This parameterization of k1(T) was used as input to astrochemical models of dark clouds, hot core/corinos, and Asymptotic Giant Star (AGB) stellar outflow environments using the UMIST Rate12 (UDfa) network of reactions. For a range of temperatures, the models yielded no significant changes in the abundances of HCN, HNC, and HCO upon inclusion of their formation using the rate coefficients reported here. It was found that the new rate coefficients for the reaction CN + CH2O do not have a significant effect on the modeled abundances of reagents and products of this reaction due to several other competing reactions with large rate coefficients. However, we note that the uncertainty at the CCSD(T)/aug-cc-pVTZ level of theory, (∼3.0–4.5 kJ mol−1) is similar to the difference in energy (4.59 kJ mol−1) calculated for the transition states forming HCN+HCO and HNC+HCO products. Hence, although HCN+HCO are the most likely products, it is possible that the HNC channel may also be the dominant one, as within the calculation errors the HNC+HCO transition state might be the lower one. In order to examine the impact of the scenario in which HNC is the dominant product of the reaction, the rate coefficient k1 was maintained at the same value, but it was assumed that HNC+HCO formed 100% of the products. For a 10 K rate coefficient and 100% formation to HNC, the title reaction provides less than 1% of the overall HNC production rate for cold environments such as molecular clouds. For the Orion Hot Core case at higher temperatures, the HNC+HCO reaction provides around 1% of the total HNC production rate. So, assuming 100% HNC production rather than very little production from the CN+CH2O reaction made a negligible difference to the modelled HNC abundance.
The main astrochemical implication of this paper is that the title reaction cannot contribute to the formation of formyl cyanide, HCOCN, in interstellar clouds, as has been suggested previously and currently implemented in astrochemical models.
Footnotes |
† Electronic supplementary information (ESI) available: The experimental method (Sections S1–S3), computational methods using Gaussian (Section S4), MESMER (Section S5) and capture theory (Section S6), parameterization of the rate coefficient (Section S7) and results of the astrochemical modelling (Section S8). See DOI: https://doi.org/10.1039/d2cp05043a |
‡ Current address: Andor Technology Ltd., Belfast, BT12 7AL, UK. |
§ Current address: School of Chemistry, University of Birmingham, Edgbaston, B15 2TT, UK. |
This journal is © the Owner Societies 2023 |