• Non ci sono risultati.

Bimolecular Homolytic Substitutions at Nitrogen: An Experimental and Theoretical Study on the Gas-Phase Reactions of Alkyl Radicals with NF3

N/A
N/A
Protected

Academic year: 2021

Condividi "Bimolecular Homolytic Substitutions at Nitrogen: An Experimental and Theoretical Study on the Gas-Phase Reactions of Alkyl Radicals with NF3"

Copied!
1
0
0

Testo completo

(1)

This is the author's final version of the contribution published as:

Paola Antoniotti, Paola Benzi, Stefano Borocci, Chiara Demaria, Maria Giordani, Felice

Grandinetti, Lorenza Operti, Roberto Rabezzana,

Bimolecular homolytic substitutions at nitrogen: an experimental and theoretical study

on the gas phase reactions of alkyl radicals with NF

3

Chemistry – A European Journal, 21(44), 2015,

15826-15834, 0.1002/chem.201501757

The publisher's version is available at:

http://onlinelibrary.wiley.com/doi/10.1002/chem.201501757/full

When citing, please refer to the published version.

Link to this full text:

http://hdl.handle.net/2318/1524250

This full text was downloaded from iris-Aperto:

https://iris.unito.it/

(2)

Bimolecular homolytic substitutions at nitrogen: an experimental

and theoretical study on the gas phase reactions of alkyl radicals

with NF

3

Paola Antoniotti,

*[a]

Paola Benzi,

[a]

Stefano Borocci,

[b]

Chiara Demaria,

[a]

Maria Giordani,

[b]

Felice

Grandinetti,

[b]

Lorenza Operti,

[a]

Roberto Rabezzana

[a]

Abstract: The X-ray irradiation of binary mixtures of alkyl iodides R-I (R = CH3, C2H5, i-C3H7) and NF3 produces R-NF2 and R-F. Based on

calculations performed at the CCSD(T), MRCI(SD+Q), G3B3, and G3 levels of theory, the former product arises from a bimolecular homolytic substitution (SH2) by the alkyl radicals R, which attack the N atom of NF3. This mechanism is consistent with the suppression of R-NF2 by

addition of O2 (an efficient alkyl-radical scavenger) to the reaction mixture. The R-F product arises from the attack of R to the F atom of NF3,

but additional contributing channels are conceivably involved. The F-atom abstraction is, indeed, considerably more exothermic than the SH2

reaction, but the involved energy barriers are comparable, and the two processes are comparably fast.

Introduction

Bimolecular homolytic substitutions (SH2) involving radical

species Ri according to Equation (1)

R1 + Y-R2  Y-R1 + R2

(1)

are elementary steps of many chemical reactions and important synthetic tools,[1] particularly for the formation of carbon-carbon

and carbon-heteroatom bonds. These processes typically occur at the univalent hydrogen or halogen atoms (Y = H, Cl, Br, I), as well as at the main-group heteroatoms (Y = Si, Ge, Sn, P, O, S, and Se).[2]

Over the years, the mechanism of reaction (1) has attracted considerable experimental and theoretical interest. A first generally accepted route is the approach of R1 along a trajectory

opposite to R2 (backside attack), passing through a co-linear (or

nearly co-linear) transition structure (TS) or a hypervalent intermediate.[3] For example, smooth TSs are likely involved in

the SH2 reactions between alkyl radicals and S or Se substrates,

Te substrates reacting by formation of hypervalent intermediates.[4] Hypervalent intermediates are also predicted in

the attack of methyl and other radicals at the P atom of methyl phosphine and phosphine oxides, and these theoretical suggestions are consistent with the experimental evidence.[5] An

alternative accepted route is the frontside attack of R1, and the

two mechanisms may be, sometimes, competitive. This occurs, for example, in the SH2 reactions of methyl and acyl radicals at

group 14 heteroatoms.[6] The conceivable occurrence of

frontside and backside mechanisms in the homolytic attack of silyl, germyl, and stannyl radicals at silicon, germanium, and tin has been also theoretically investigated by Schiesser and co-workers.[7] On the other hand, the mechanism of S

H2 reactions

between radical species and nitrogen molecules is, still, essentially unexplored. Processes like these were first reported in 1978 by Cadman et al.,[8] who analyzed the products of the

reactions of the alkyl radicals R = CH3, CF3, C2H5, i-C3H7, and

t-C4H9 (generated by UV photolysis of ketones) with NF3, and

observed the formation of the corresponding R-NF2. The

F-abstraction products R-F were, instead, detected only for R = i-C3H7 and t-C4H9. More recently, searching for a method for the

direct and continuous generation of N,N–difluoroamines, Belter et al.[9] explored the reactivity of NF

3 with aliphatic and aromatic

substrates. They found that certain alkanes, cycloalkanes, and ethers reacted with NF3 at high temperature to produce the

corresponding difluoroamine. Little amounts of direct fluorination were also observed. The experimental observation of SH2

reactions involving NF3 raises stimulating questions concerning

their detailed mechanisms. Our interest in this subject is also driven by the attention that we recently focused on the gas-phase ionic reactions occurring in mixtures of group 14 hydrides and fluorinated compounds, including NF3.[10] Therefore, we

decided to experimentally investigate the reactions between NF3

and the exemplary radicals CH3, C2H5, and i-C3H7, generated by

X-ray irradiation of the corresponding iodides R-I, and to perform

ab initio and density functional theory (DFT) calculations on the

observed reactions. The results of our investigation will be discussed in the present article.

Results and Discussion

Radiolysis of R-I/NF3 mixtures (R = CH3, C2H5, i-C3H7)

The X-ray irradiation of R-I/NF3 mixtures is expected to produce

both radicals and ions, and their relative contribution to the formed products is evaluated in terms of the average energy absorbed to form the ion pair, W,[11,12] and the ionization energy,

I of the various precursors. The difference between W and I is,

in particular, always positive, and the W-I excess energy is available to form excited molecules and/or radicals. The most abundant radicals arise, in particular, from the dissociations [a] Dr. P. Antoniotti*, Dr. P. Benzi, Dr. C. Demaria, Prof. L. Operti, Prof.

R. Rabezzana Dipartimento di Chimica Università di Torino

Via P. Giuria 7, 10125 Torino (Italy) E-mail: paola.antoniotti@unito.it

[b] Dr. S. Borocci, Dr. M. Giordani, Prof. F. Grandinetti

Dipartimento per la Innovazione nei sistemi Biologici, Agroalimentari e Forestali (DIBAF)

Università della Tuscia

L.go dell’Università, s.n.c., 01100 Viterbo (Italy)

and

Istituto di Metodologie Chimiche, Area della Ricerca di Roma 1 Via Salaria Km 29,300, 00015 Monterotondo (Italy)

(3)

NF3  NF2 + F

(2)

R-I  R + I (3) (R = CH3, C2H5, i-C3H7)

The W and I values of NF3 and R-I are listed in Table 1 together

with the enthalpy changes of reactions (2) and (3) derived from experimental thermochemical data.[13] It is thus possible to

estimate that the radicals formed for each ion are nearly seven from NF3, nearly six from CH3I, and nearly five from both C2H5I

and i-C3H7I. Thus, radical reactions are expected to give the

major contribution to the observed products.

Table 1. Mean energy absorbed to form an ion pair W [J molecule-1], ionization

potential I [J molecule-1], and dissociation enthalpy ΔH at 298.15 K [J molecule -1] of NF

3 and R-I.

[a] The W of CH

3I is the average value taken from Ref. [43]. The experimental

values of NF3, C2H5I, and i-C3H7I are not available, but it is known that, for

gaseous molecules, W/I ranges from 2.2 to 2.6.[44] Thus, W was inferred from

the average value of W/I. [b] Taken from Ref. [13].[c] Referred to the homolytic

dissociation of the F2N-F and R-I bonds, and calculated using the

thermochemical data quoted in Ref. [13].

The gaschromatography-mass spectrometry (GC-MS) analysis of the irradiated R-I/NF3 mixtures unravelled the formation of

appreciable amounts of the gaseous R-NF2 and R-F, together

with small amounts of additional products (for example, the irradiation of CH3I/NF3 mixtures produced small amounts of ethyl

iodide and ethylenediamine). It is useful to remember that the fraction of the total energy absorbed by each species is proportional to its mass. The formation of new products from radiolysis of R-X/NF3 mixtures change the gas phase

composition and causes a different sharing of the absorbed energy among the various gaseous species. Therefore the yields of the observed final products vary in a not predictable way and an irregular trend of the gas species amounts can also be observed comparing the results obtained with different irradiation dose. The products detected from experiments performed with two different irradiation doses (5 kGy and 50 kGy), in the absence/presence of O2 are, in particular, reported in Table 2. In

the irradiation of CH3I/NF3 mixtures, the relative amount of

CH3NF2 and CH3F is nearly independent on the irradiation dose.

However, in the presence of O2 (an efficient scavenger of alkyl

radicals), CH3F tends to decrease, but CH3NF2 almost

disappears.

Table 2. -moles of R-NF2 and F obtained from the X-ray irradiation of

R-I/NF3 and R-I/NF3/O2 mixtures (R = CH3, C2H5, i-C3H7) for different irradiation

doses [kGy]. a CH3I/NF3 CH3I/NF3/O2 5 50 5 50 CH3F 10.1 550 8.12 204 CH3NF2 4.40 231 0.10 3.29 CH3F/CH3NF2 2.30 2.38 81.2 62.0 C2H5I/NF3 C2H5I/NF3/O2 C2H5F 29.0 688 16.1 C2H5NF2 32.2 471 n.d. C2H5F/C2H5NF2 0.90 1.46 -i-C3H7/NF3 i-C3H7/NF3/O2 C3H6 262 1280 170 i-C3H7F 25.2 591 141 i-C3H7NF2 4.24 68.1 n.d. i-C3H7F /i-C3H7NF2 5.94 8.68 -a Determinations are affected by an error of ca. ±15%.

This suggests that the latter product mainly arises from the reaction between CH3 radicals and NF3. This process, however,

only partially contributes to the formation of CH3F, and additional

forming channels must be considered (vide infra). The results of additional experiments performed on mixtures with different CH3I/NF3 partial pressures (see Table 3) confirm that the

formation of CH3NF2 occurs from the methyl radical through a

mechanism of substitution to the nitrogen atom of NF3.In fact,

the amount of CH3NF2 increases in the mixtures containing CH3I

at the pressure of 300 Torr, and invariably disappears in the presence of O2. As for CH3F, its detected amount increases by

increasing the relative pressure of CH3I, and a similar

dependence is obtained in the presence of O2. This indicates

that it does not arise exclusively from the reaction of methyl radicals with NF3. We also note that, at the same relative

pressure of CH3I, the rise of NF3 increases the formed CH3F by

nearly 25%, and a similar trend is observed in the presence of O2. Moreover, considering the amounts of CH3F which are

formed at various relative pressure of the reactants, this product decreases by nearly 25% by addition of O2. This allows to

estimate that the attack of CH3 to the F atoms of NF3 contributes

as nearly 25% of the product formation.

Table 3. µ-moles of R-NF2 and R-F obtained from X-ray irradiation of CH3I/NF3

and CH3I/NF3/O2 mixtures for different CH3I/NF3 partial pressures [Torr] at 5

kGy irradiation dose. a

CH3I/NF3 CH3I/NF3/O2 150/300 300/30 0 300/150 150/300/110 300/300/110 300/150/110 CH3F 8.93 19.20 15.29 6.59 14.23 11.30 CH3NF2 4.69 25.90 9.76 n.d. n.d. n.d. CH3F/ CH3NF2 2.15 0.74 1.57 - -

-a Determinations are affected by an error of ca. ±15%.

The products obtained from the irradiation of C2H5I/NF3 mixtures

are C2H5NF2 and C2H5F, their yields ratio approximately

increasing by 1.5 by increasing the irradiation dose. The addition of O2 causes again the disappearance of C2H5NF2, C2H5F only

decreasing. This confirms that the former product arises from reactions involving the alkyl radicals. The irradiation of i-C3H7I/NF3 mixtures produces i-C3H7NF2 and i-C3H7F, together

with large amounts of propylene. This additional product, however, does not contribute to the formation of the former species. In fact, no products were detected from the irradiation of propylene/NF3 and propylene/NF3/O2 mixtures. In any case,

compared with the R-F/R-NF2 yield ratios detected from the

irradiated R-I/NF3 (R = CH3, C2H5), the i-C3H7F/i-C3H7NF2 yield

ratio is significantly different. The alkylfluoride is, in particular, much more abundant than the alkyldifluoroamine, the ratio being nearly 6, and becoming even higher by increasing the irradiation dose. This suggests that i-C3H7F likely arises from

mechanisms different from those responsible of the formation of CH3F and C2H5F. The addition of O2 causes an appreciable

3

W[a] I[b] W - I ΔH[c] radical per ion

NF3 5.18×10-18 2.16×10-18 3.02×10 4.03×10-1 7.5

CH3I 4.14×10-18 1.52×10-18 2.62×10 3.94×10-1 6.6

C2H5I 3.60×10-18 1.50×10-18 2.10×10 3.86×10-1 5.4

(4)

increase of i-C3H7F, further confirming that it is obtained through

various reactions. In contrast, i-C3H7NF2 completely disappears,

thus confirming its origin from reactions involving isopropyl radicals and NF3.

Computational results

The experiments suggest that the R-NF2 obtained from the X-ray

irradiation of R-I/NF3 mixtures (R = CH3, C2H5, i-C3H7) arise from

the bimolecular homolytic substitution (4) occurring at the nitrogen atom. The concomitant detection of the corresponding

R-F suggests the competitive occurence of the F-atom

abstraction (5), involving the attack of R to the F atom(s) (even though additional channels contributing to this product must be assumed): R + NF3  R-NF2 + F (4) R + NF3  R-F + NF2 (5) (R = CH3, C2H5, i-C3H7)

Both these reactions were investigated by theoretical calculations, and found to proceed through the intermediates and TSs whose CASSCF/6-31G(d) and B3LYP/6-31G(d) optimized geometries are shown in Figures 1, 3, and 5. The corresponding potential enthalpy diagrams obtained at the CCSD(T)/6-311G(2d,2p)//CASSCF(9,6)/6-31G(d) level of theory are reported in Figures 2, 4 and 6. The energy (E), enthalpy (H), and free energy (G) differences of the various species, computed at the CCSD(T)/6-311G(2d,2p)//CASSCF/6-31G(d), at the MRCI(SD+Q)/6-311G(d,p)/CASSCF/6-31G(d), and at the G3B3 levels of theory, are listed in Tables 4, 5 and 6. Also included in Table 4 are the E, H, and G of the species involved in the reaction between CH3 and NF3 as computed at

the G3 level of theory.

Reactions between CH3 radical and NF3

Based on experimental thermochemical data,[14,15] for R = CH 3,

reaction (4) is slightly exothermic by 2.2 kcal mol-1 (the value

quoted so far by Cadman et al.[8] was 1.8 kcal mol-1), and

reaction (5) is definitely exothermic by 49.2 kcal mol-1.[14] These

data are consistent with the theoretical predictions. Thus (see Table 4), at the CCSD(T), MRCI(SD+Q), G3B3, and G3 levels of theory, reaction (4) is exothermic by 10.5, 12.2, 7.0, and 6.6 kcal mol-1, respectively, and reaction (5) is exothermic by 51.8, 50.7,

51.7, and 51.8 kcal mol-1, respectively. It is well known from

previous studies[16-19] that, in the gaseous phase, NF 3 is a

bifunctional Lewis base, able to interact with electrophilic species by both the N- and the F atom(s). We, therefore, explored the attack of CH3 to both the N and the F atom(s) of

NF3, and actually located the two energy minima 1 and 2, and

the three TSs TS_1, TS_2, and TS_3 shown in Figure 1. These minima and TSs are connected as shown in Figure 2. TS_1 (imaginary frequency: 1108.8i cm-1) is the TS for the homolytic

substitution of CH3 at the N atom of NF3 with extrusion of F, and

TS_2 (imaginary frequency: 1161.6i cm-1) is the TS for the

F-atom abstraction. As shown in Figure 1, TS_1 features an expectedly-long attacking C-N bond distance, predicted, in particular, as 2.101 Å, 2.072 Å and 1.886 at the CASSCF,

B3LYP and MP2(full) level, respectively. The predicted N-F distance is 1.746 Å, 1.595 Å and 1.496 Å, respectively, at the CASSCF, B3LYP and MP2(full) levels of theory.

The C-N-F angle results as 153.9 at the CASSCF, and comparable values of 154.0 and 157.2°, respectively, are obtained at the B3LYP and MP2(full) levels. At any computational level, TS_2 adopts a nearly co-linear arrangement of the entering CH3 and the leaving NF2. At the CASSCF and

B3LYP levels of theory, the C-F and N-F bond lengths are comparable and amount, respectively, to 2.083 Å and 1.731 Å,

and 2.079 Å and 1.612 Å. The MP2(full) furnishes,

Figure 1. CASSCF(9,6)/6-31G(d) and B3LYP/6-31G(d) (italics) optimized

geometries (Å and °) of the species involved in the reactions between CH3

and NF3.

Figure 2. CCSD(T)/6-311G(2d,2p)//CASSCF(9,6)/6-31G(d) relative enthalpies

at 298.15 K (kcal mol-1) of the species involved in the reactions between CH 3

and NF3.

instead, lower values of 1.862 Å and 1.553 Å. Interestingly, at any computational level, for both TS_1 and TS_2 the bond distance involving the entering group is long, the distance involving the leaving group being slightly shorter using the B3LYP method, and further decreasing at the MP2(full) level of theory. The CCSD T1diagnostic of TS_1, 0.039, and TS_2, 0.037, are higher than the accepted threshold of 0.020, and this suggests a contribution of multi-determinantal character to their wave-function. However, for both species, the CI coefficient of the most important contribution to the ground-state CASSCF wave function resulted as 0.93. This indicates the by far predominance of a single configuration, and suggests also the prevailing role of dynamical rather than static correlation. The comparison between the CCSD and MRCI(SD+Q) activation

(5)

enthalpies further confirmed the minor role of multi-reference effects in the investigated systems. Thus, the enthalpy barrier of the SH2 reaction (4) passing through TS_1 is 21.9 kcal mol-1 at

the CCSD(T) level, and 23.3 kcal mol-1 at the MRCI(SD+Q), and

the enthalpy barrier of the F-atom abstraction (5), passing

through TS_2, is 17.8 kcal mol-1 at the CCSD(T) level, and 18.8

kcal mol-1 at the MRCI(SD+Q). The two barriers are also

predicted as 19.4 kcal mol-1 and 17.5 kcal mol-1, respectively, at

the G3B3 level of theory, and 20.9 kcal mol-1 and 19.8 kcal mol-1

at the G3 level of theory (see Table 4). Thus, the enthalpy

Table 4. Relative energies ∆E at 0 K [kcal mol-1], relative enthalpies ∆H at 298.15 K [kcal mol-1], and relative free energies ∆G at 298.15 K [kcal mol-1] of the

species involved in the reactions between CH3 and NF3 (see also Figure 1).

Species CCSD(T)/6-311G(2d,2p)[a] MRCI(SD+Q)/6-311G(d,p)[a] G3B3[b] G3

∆E ∆H ∆G ∆E ∆H ∆G ∆E ∆H ∆G ∆E ∆H ∆G

CH3 + NF3 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1 -11.1 -11.7 -3.2 -12.6 -13.3 -4.8 -12.5 -13.5 -4.6 -10.0 -10.5 -2.7 2 -54.3 -54.2 -47.7 -52.5 -52.4 -45.9 -52.3 -52.2 -47.4 -52.4 -52.3 -46.1 TS_1 22.7 21.9 32.4 24.1 23.3 33.8 20.3 19.4 28.5 22.1 20.9 32.4 TS_2 18.2 17.8 25.8 19.2 18.8 26.8 17.9 17.5 24.1 20.4 19.8 29.1 TS_3 37.5 37.1 46.0 35.4 35.1 44.0 - - - -CH3NF2 + F -9.5 -10.5 -1.7 -11.2 -12.2 -3.4 -6.3 -7.0 -5.1 -5.9 -6.6 -3.0 CH3F + NF2 -51.9 -51.8 -48.7 -50.8 -50.7 -47.6 -51.3 -51.7 -52.8 -51.4 -51.8 -51.8

[a] At the CASSCF/6-31G(d) optimized geometries. [b] At the B3LYP/6-31G(d) optimized geometries. [c] At the MP2(full)/6-31G(d) optimized geometries.

difference between TS_1 and TS_2 results as 4.1 and 4.5 kcal mol-1, respectively, at the CCSD(T) and MRCI(SD+Q) levels, it

decreases to only 1.9 kcal mol-1 at the G3B3 level, and further

reduces to 1.1 kcal mol-1 at the G3 level. These differences likely

reflect the different CASSCF, MP2, and B3LYP geometries. According to the IRC calculations, TS_1 does not directly evolve into CH3NF2 + F. Rather, it connects the reactants with the

complex 1, which features (see Figure 1) a F atom relatively distant from the N atom of CH3NF2. The predicted N-F bond

length depends on the computational level, and, in particular, progressively reduces from the CASSCF (2.943 Å) to the MP2(full) (2.419 Å), and the B3LYP level (1.981 Å). The eventual dissociation of 1 requires to overcome a small enthalpy barrier, predicted as nearly 1 kcal mol-1 at the CCSD(T) and

MRCI(SD+Q) levels of theory, and slightly higher at the G3 (3.9 kcal mol-1), and G3B3 levels of theory (6.5 kcal mol-1). TS_2

connects the reactants with the weakly-bound complex 2 (see Figure 2), which arises from the interaction of the F atom of CH3F with the N atom of NF2. It features a rather long N-F bond

distance, predicted, in particular, as 2.902 Å, 2.808 Å and 2.869 Å, respectively, at the CASSCF, at the MP2(full), and at the B3LYP levels of theory. Consistently, the dissociation enthalpy is only slightly positive at the CASSCF (2.4 kcal mol-1), and at the

MRCI(SD+Q) level of theory (1.7 kcal mol-1), and even lower (0.5

kcal mol-1) at the G3 and G3B3 levels, actually disappearing by

inclusion of the entropy contribution. As shown in Figure 2, complex 2 is also accessible from complex 1 by overcoming the high enthalpy barrier [48.8 kcal mol-1 at the CCSD(T), and 48.4

kcal mol-1 at the MRCI(SD+Q) level of theory] corresponding to

TS_3 (imaginary frequency: 855.6i cm-1). According to the IRC

calculations, the process consists of the insertion of a F atom into the C-N bond of 1. The CCSD T1 diagnostic of TS_3 (0.038), and the highest CI coefficient contributing to the CASSCF wave function (0.85) indicate the relatively-high contribution of the static correlation. A TS analogous to TS_3 (imaginary frequency: 874.3i cm-1) lying 29.8 kcal mol-1 above

the reactants was also located at the B3LYP level of theory, but the IRC calculations revealed that it connects 2 with the separated H2CNF, HF, and F. Thus, it was not further explored

at the G3B3 level of theory.

Reactions between C2H3 and NF3

Based on the experimental thermochemical data, for R = C2H5,

reaction (4) is slightly exothermic (-2.36 ± 0.5 kcal mol-1 [14,15]) or

nearly thermoneutral (0.2 kcal mol-1 [8]), and reaction (5) is

strongly exothermic (48.3 kcal mol-1[8] and 52.5 ± 0.5 kcal mol-1 [14]). The theoretical study of these processes revealed

mechanisms strictly similar to those outlined for the

corresponding reactions involving CH3.Thus, C2H5 attacks the N

atom of NF3 and passes through TS_1' (imaginary frequency:

1090.8i cm-1) so to form the weakly-bound complex 1', which, in

(6)

turn, dissociates into C2H5NF2 and F. The dissociation enthalpy

Figure 3. CASSCF(9,6)/6-31G(d) and B3LYP/6-31G(d) (italics) optimized

geometries (Å and °) of the species involved in the reactions between C2H5

and NF3.

Figure 4. CCSD(T)/6-311G(2d,2p)//CASSCF(9,6)/6-31G(d) relative enthalpies

at 298.15 K (kcal mol-1) of the species involved in the reactions between C 2H5

and NF3.

amounts to 2 kcal mol-1 at the CCSD(T) and MRCI(SD+Q)

levels of theory, and increases to 4.4 and 7.4 kcal mol-1 at the

G3 and G3B3 levels, respectively. These differences parallel the different C2H5N(F2)---F bond lengths predicted at the CASSCF,

MP2(full), and B3LYP levels of theory. With respect to the reagents, 1' is more stable by 11.6, 13.5 and 15.1 kcal mol-1,

Table 5. Relative energies at 0 K (kcal mol-1), enthalpies at 298.15 K (kcal mol-1) and free energies at 298.15 (kcal mol-1) of the species involved in the reactions

between C2H5 and NF3 (see Figure 3 ).

Species CCSD(T)/6-311G(2d,2p)[a] MRCI(SD+Q)/6-311G(d,p)[a] G3B3[b]

∆E ∆H ∆G ∆E ∆H ∆G ∆E ∆H ∆G

C2H5 + NF3 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1' -12.9 -11.6 -15.4 -14.8 -13.5 -17.3 -14.5 -15.1 -4.1 2' -58.4 -56.5 -62.1 -56.5 -54.5 -60.1 -56.0 -55.8 -48.2 TS_1' 19.6 20.9 18.4 20.2 21.5 19.0 16.9 16.4 27.2 TS_2' 14.7 15.7 13.6 16.6 17.5 15.5 14.6 14.5 22.8 TS_3' 29.9 31.4 28.2 38.6 40.2 36.9 - - -CH3NF2 + F -11.2 -9.7 -17.0 -13.0 -11.5 -18.8 -7.3 -7.7 -3.7 CH3F + NF2 -55.6 -53.7 -63.3 -54.4 -52.6 -62.2 -54.5 -54.9 -54.6

(7)

respectively, at the CCSD(T), MRCI(SD+Q), and G3B3 levels of theory, and TS_1' is less stable by 20.9, 21.5, 16.4, and 19.2 kcal mol-1, respectively, at the CCSD(T), MRCI(SD+Q), G3B3, and G3 levels of theory. Thus, the activation barrier

corresponding to TS_1' is comparable with that corresponding to TS_1, the difference amounting to only 1.0 and 1.8 kcal mol-1

at the CCSD(T) and MRCI(SD+Q) levels of theory, and only slightly increasing at the G3B3 level (3.0 kcal mol-1). In addition, the

comparative examination of Figures 1 and 3 reveals that the geometrical parameters of 1' and TS_1', and those of the corresponding analogues 1 and TS_1 are strictly similar. Consistent with the experimental thermochemical data, the F-abstraction between C2H5 and NF3 by is predicted to be exothermic by 53.7, 52.6, and 54.9 kcal mol-1 at the CCSD(T),

MRCI(SD+Q), and G3B3 levels, respectively. The reaction passes through TS_2' (imaginary frequency: 1191.2i cm-1), and proceeds

with formation of complex 2', which, in turn, undergoes the only slightly endothermic loss of NF2 [the enthalpy change is 2.8 and

1.9 kcal mol-1 at the CCSD(T) and MRCI(SD+Q) levels,

respectively, and decreases to only 0.9 kcal mol-1 at the G3B3

level, disappearing by inclusion of the entropy contribution]. The activation barrier corresponding to TS_2' is 15.7, 17.5, and 14.5 kcal mol-1 at the CCSD(T), MRCI(SD+Q), and G3B3 levels of

theory, respectively, and these values are slightly lower (by ca. 1-3 kcal mol-1) than those predicted for the analogous reaction

involving CH3. Complexes 1' and 2' are connected through the

high energy TS_3' (imaginary frequency: 885.3i cm-1), which is the

corresponding analogue of TS_3. Its T1 diagnostic (0.040), and highest CI coefficient contributing to the CASSCF wave function

(0.86) indicate the relatively-high contribution of static correlation. As a matter of fact, this TS was not located at the B3LYP level of theory. We finally note that TS_2', TS_3', and 2' are structurally similar to their corresponding analogues TS_2, TS_3, and 2.

Reactions between i-C3H7 and NF3.

Based on experimental thermochemical data, for R = i-C3H7,

reaction (4) is weakly exothermic (-4.66 ± 0.5 kcal mol-1 [14,15]) or

nearly thermoneutral (-0.1 kcal mol-1 [8]), and reaction (5) is strongly

exothermic (-46.1 kcal mol-1 [8] or -50.3 ± 0.5 kcal mol-1 [14,15]).

According to the calculations, the attack of i-C3H7 to NF3 may occur

to both the N and the F atom(s), and reactions (4) and (5) proceed by mechanisms strictly similar to those outlined for CH3 and C2H5.

We located, in fact, both TS_1" and TS_2", which are, respectively, the corresponding analogues of TS_1 and TS_1', and

TS_2 and TS_2'. They are also structurally quite similar to their

congeners (see Figure 5). Based on Table 6, the enthalpy barrier of the SH2 reaction (5) (passing through TS_1") is 14.5 kcal mol-1

at the CCSD(T) level, and 13.5 kcal mol-1 at the G3B3 level. At the

same levels, the enthalpy barrier of the F-abstraction (5) (passing through TS_2") amounts, respectively, to 9.7 kcal mol-1 and

11.2 kcal mol-1. The inspection of Tables 4, 5, and 6 reveals that the activation enthalpies corresponding to TS_1" and TS_2"

are both lower than those corresponding to TS_1 and TS_1', and TS_2 and TS_2' by nearly 7 kcal mol-1 at the CCSD(T) level of

theory.

Figure 5. CASSCF(9,6)/6-31G(d) and B3LYP/6-31G(d) (italics) optimized geometries (Å and °) of the species involved in the reactions between i-C3H7 and

NF3.

Figure 6. CCSD(T)/6-311G(2d,2p)//CASSCF(9,6)/6-31G(d) relative enthalpies at 298.15 K (kcal mol-1) of the species involved in the reactions between

i-C3H7 and NF3.

In addition, at the G3B3 level of theory, the activation enthalpies decrease by nearly 6 kcal mol-1 passing from TS_1 (19.4 kcal

mol-1) to TS_1" (13.5 kcal mol-1), and from TS_2 (17.5 kcal mol-1) to TS_2" (11.2 kcal mol-1). It is also of interest to note the strict

similarity of the CCSD(T) enthalpy differences between TS_1" and TS_2" (4.8 kcal mol-1), TS_1 and TS_2 (4.1 kcal mol-1), and

TS_1' and TS_2' (5.2 kcal mol-1).

Comparison between the experimental and the theoretical data

The theoretical results suggest that, despite their largely different exothermicities, for any radical R (R = CH3, C2H5, i-C3H7), the

energy barriers of the SH2 reaction (4), and of the F-abstraction (5) are, essentially, comparable. The TS of reaction (4) is, in

fact, predicted to be more stable than that of reaction (5), but the differences range between 4 and 5 kcal mol-1 at CCSD(T) level

of theory, and further reduce to nearly 2 kcal mol-1 at G3B3 and G3 levels. These values are, indeed, well within the combined

uncertainties of the employed methods, expected to be ca. ± 2-4 kcal mol-1. These results together with the fundamentally

monodeterminantal character of the wave function for the transition structures for both reaction mechanisms also enable us to

iris-AperTO

University of Turin’s Institutional Research Information System and Open Access Institutional

Repository

(8)

assert that the less expensive G3B3 method is suitable to describe the two reaction mechanisms considered. In any case, it is possible to note that the absolute barriers depend on the radical and, in particular, decrease in the order i-C3H7 > C2H5 ≥ CH3.

Table 6. Relative energies at 0 K (kcal mol-1), enthalpies at 298.15 K (kcal mol-1) and free energies at 298.15 K (kcal mol-1) of the species involved in the

reactions between i-C3H7 and NF3 (see Figure 5).

Species CCSD(T)/6-311G(2d,2p)[a] G3B3 ∆E ∆H ∆G ∆E ∆H ∆G i-C3H7 + NF3 0.0 0.0 0.0 0.0 0.0 0.0 1" -16.3 -16.4 -6.0 -16.9 -17.5 -5.3 2" -62.3 -61.9 -53.5 -58.7 -58.5 -49.8 TS_1" 14.6 14.5 25.9 13.8 13.5 24.7 TS_2" 8.2 9.7 13.7 11.0 11.2 19.5 TS_3" 31.6 32.0 41.1 29.3 28.8 40.7 CH3NF2 + F -14.1 -15.2 -5.1 -9.2 -9.5 -4.9 CH3F + NF2 -59.4 -59.6 -53.2 -57.3 -57.6 -56.8

[a] At the CAS-MCSCF/6-31G(d) optimized geometries. [b] At the B3LYP/6-31G(d) optimized geometries.

The theoretical suggestion of comparable energy barriers is consistent with the experiments, which show, in particular, that the X-ray irradiation of R-I/NF3 mixtures (R = CH3, C2H5, i-C3H7) produces comparable amounts of R-NF2 and R-F. The origin of the

former product by a reaction involving the radical R is confirmed by the fact that, in the presence of O2 (an efficient scavenger of

alkyl radicals), its yield dramatically decreases. The formation of R-F is also reduced, but not suppressed, by the presence of O2, and this suggests the occurrence of forming channels other than reaction (5). A conceivable contribution to the observed

residual methyl fluoride is the ionic reaction between CH3+ and NF3. Thus, in a previous study[10c] performed by ion trap mass

spectrometry, we found that, in ionized CH4/NF3 mixtures, the main product from the efficient reaction between CH3+ and NF3

was CH3F, arising from the attack of the methyl cation to the fluorine atom of NF3, and subsequent dissociation. We also do not

rule out the occurrence of a reaction between F atoms (not captured by O2) and CH3I. We would also expect, however, the

competitive occurrence of a H-atom abstraction by F radicals so to form HF and CH2I.[20] In fact, hydrogen abstraction reactions

by F radicals from alkyl halides or alkanes are highly exothermic, and play a key role, for example, in the etching processes in plasma and CVD reactors.[21]

Conclusions

The SH2 reactions between radical species and nitrogen compounds are, still, essentially unexplored. The present X-ray

experiments and high-level theoretical calculations indicate that processes like these are involved in the reactions between alkyl radicals and NF3. These processes are considerably less exothermic than the competitive F-atom abstractions (with formation of

R-F), but the involved barriers are comparably high, and the two reactions are comparably fast. Put in the perspective, the

simplest fluoroamines appear as useful exemplary substrates to further investigate SH2 reactions at nitrogen. It will be certainly

of interest to assay the conceivable occurrence of these processes by radicals containing the heaviest elements of the group 14, as well as other heteroatoms. We are planning work along this directions.

Experimental Section

X-ray radiolysis

NF3 at 99.99% stated purity, and pure O2 were supplied by Praxair and SIAD, respectively. Methyl, ethyl, and

isopropyl iodides, best grade, were supplied by Sigma Aldrich, and purified as described previously.[22] The R-I/NF

3 mixtures (R

= CH3, C2H5, i-C3H7), with 165 Torr of R-I and 360 Torr of NF3 partial pressures (1 Torr = 133 Pa), were prepared in 350 ml

Pyrex vessels. Two sets of samples were irradiated at absorbed irradiation doses of 5 and 50 kGy: the first one consisted of R-I/NF3 mixtures, and the second one consisted of R-I/NF3 mixtures, with 110 Torr of O2 added as radical scavenger. For the CH3I/

NF3 mixture, experiments were also performed at the variable relative pressures (Torr) of 150/300, 300/300, and 300/150. The

absorbed irradiation dose was 5 kGy. All the samples were irradiated using a CPXT-320 tube (GILARDONI) as the X-ray source with a maximum output of 320 KeV. Standard vacuum techniques were used to handle reactants and gaseous products.

Gas Chromatography - Mass Spectrometry analyses

After irradiation, a small amount of the gaseous phase was collected for qualitative and quantitative analyses of volatile compounds by GC-MS. A Varian 3400/ Finnigan ITD instrument was employed, equipped with a Alltech AT-1 capillary column (polydimethylsiloxane, 30 m long, 0.25 mm internal diameter, 1.0 mm film thickness). Before injection, the GC oven was cooled at 193 K by introducing liquid nitrogen; afterwards the column was heated up to 433 K with the following temperature program: isothermal starting step at 193 K for 4 minutes; heating step up to 373 K (20 K/min); isothermal step at 373 K for 3 minutes; heating step up to 433 K (20 K/min); isothermal step at 433 K for 3 minutes; cooling step to room temperature. Split of about 16 ml min-1 was applied during injection; helium was used as carrier gas with 0.8 ml min-1 flow. Electron ionization was performed at

70 eV, and the spectra were collected in the 15 - 500 u mass range.

iris-AperTO

(9)

Computational methods

The calculations were performed with the GAUSSIAN09[23] and MOLPRO 2010.1[24] programs. The geometries of the reagents,

intermediates, products, and TSs involved in the reactions between the CnH2n+1 radicals (n = 1-3) and NF3 were fully-optimized at

the complete active space multi-configuration self-consistent field level of theory[25-27] in conjunction with the 6-31G(d)

basis set[28] [CASSCF/6-31G(d)] by gradient-based techniques[29-32] and with no symmetry constraints. The CASSCF

wave-function, labelled as (9,6), was built up by distributing nine electrons in the six orbitals which are most reasonably involved in the reaction mechanisms. With reference to the reactants, we included, in particular, the singly-occupied sp hybrid orbitals of the C atom of the CnH2n+1 radicals, and five orbitals of NF3, namely a pair of bonding and antibonding N-F sigma orbitals

(N-F and *N-F), two p orbitals of F, and the n orbital of N. Any located critical point was unambiguously characterized as an

energy minimum or a TS by calculating its analytical or numerical harmonic vibrational frequencies. Any TS was also related to its interconnected energy minima by intrinsic reaction coordinate (IRC) calculations.[33] The CASSCF/6-31G(d) unscaled

frequencies were also used to calculate the zero-point vibrational energies (ZPE), and the vibrational contribution to the thermal correction (TC), obtained at 298.15 K by standard statistical mechanics formulas.[34] The overall TC term was finally

obtained by adding the translational (3/2 RT) and rotational (RT or 3/2 RT) contributions at this temperature.

Total entropies were also obtained by CASSCF unscaled frequencies and moments of inertia. The absolute energies were refined by performing, at the CASSCF optimized geometries, single-point calculations at the coupled cluster level of theory[35,36]

using the 6-311G(2d,2p) basis set[28] [CCSD(T,full)/6-311G(2d,2p)], and at the multireference-configuration interaction

[MRCI(SD+Q)] computational level, as implemented in MOLPRO,[37,38] using the 6-311G(d,p) basis set

[MRCI(SD+Q)/6-311G(d,p)]. The role of higher excitations was estimated using the multireference analogue of the Davidson correction (denoted as +Q[39]), and internal contraction was used to limit the size of the calculations. The reference configurations arose from the

CASSCF(9,6) wave function, and, in order to save computational time, only the occupations with absolute coefficients in the wave function higher than 0.02 were included. The T1 diagnostics[40] were calculated at the CCSD(T,full)/6-311G(2d,2p) level of

theory. Absolute energies were also calculated using the G3[41] and G3B3[42] composite methods.

Acknowledgements

The authors thank the Università di Torino, the Università della Tuscia, and the Italian Ministero dell’Istruzione, dell’Università e della Ricerca (MIUR) for financial support.

Keywords: alkyl radicals • bimolecular homolytic substitutions • nitrogen trifluoride • theoretical calculations • x-ray radiolysis

1] a) A. G. Davies, B. P. Roberts, in J.K. Kochi (Ed.), Free Radicals, Vol.1 Wiley, New York, 1973 (Chapter 10). b) J. Fossey, D. Lefort, J. Sorba, Free Radicals in Organic Chemistry, Wiley, Chichester, 1995, p.123.

[2] a) C. H. Schiesser. L. M. Wild, Tetrahedron 1996, 42, 13265-13314; b) C. H. Schiesser. B. A. Smart, Tetrahedron 1995, 51, 3227-3338.

[3] a) C. H. Schiesser, B. A. Smart, Tetrahedron, 1995, 51, 6051-6060; b) B.A. Smart, C.H. Schiesser, J. Comput. Chem. 1995,

16, 1055

[4] a) S. M. Horvat, C. H. Schiesser, New J. Chem. 2010, 34, 1692-1699; b) C. H. Schiesser. L. M. Wild, J. Org. Chem. 1999,

64, 1131-1139.

[5] a) C. H. Schiesser. L. M. Wild, Aust. J. Chem. 1995, 48, 175-184; b) C. J. Cramer, J. Am. Chem. Soc. 1990, 112, 7965-7972; c) C. J. Cramer, J. Am. Chem. Soc. 1991, 113, 2439-2447.

[6] a) H. Matsubara, S. M. Horvat, C. H. Schiesser, Org. Biomol. Chem. 2003, 1, 1199-1203; b) H. Matsubara, C. H. Schiesser,

Org. Biomol. Chem. 2003, 1, 4335-4341.

[7] S. M. Horvat, C. H. Schiesser, L. M. Wild, Organometallics 2000, 19 (7), 1239-1246.

[8] P. Cadman, Y. Inel, A. F. Trotman-Dickenson, J. Chem. Soc., Faraday Trans. (II) 1978, 74, 2301-2312.

[9] a) R. K. Belter, J. Fluorine Chem. 2011, 132, 318-322; b) R. K. Belter, J. Fluorine Chem. 2011, 132, 961-964; c) R. K: Belter, C. A. McFerrin, J. Fluorine Chem. 2012, 135, 272-277; d) R. K. Belter, J. Fluorine Chem. 2012, 137, 73-76.

[10] a) P. Antoniotti, L. Operti, R. Rabezzana, F. Turco, G. A. Vaglio, Int. J. Mass Spectrom. 2006, 255-256, 225-231; b) P. Antoniotti, R. Rabezzana, F. Turco, S. Borocci, M. Giordani, F. Grandinetti, J. Mass Spectrom. 2008, 43, 1320-1333; c) P. Antoniotti, L. Operti, R. Rabezzana, F. Turco, C. Zanzottera, M. Giordani, F. Grandinetti, J. Mass Spectrom. 2009, 44, 1348-1358; d) P. Antoniotti, L. Operti, R. Rabezzana, F. Turco, S. Borocci, F. Grandinetti., Eur. J. Mass Spectr. 2009, 15, 209-220; e) P. Antoniotti, E. Bottizzo, S. Borocci, M. Giordani, F. Grandinetti, J. Comput. Chem. 2012, 33, 1918-1926.

[11] J. W. T. Spinks, R. J. Wood, in An Introduction to Radiation Chemistry 3rd Ed., John Wiley & Sons, New York, 1990, pp. 3-13.

[12] J. Fraden, in AIP Handbook of Modern Sensors, Physics, Designs, and Applications 4th Ed., Springer Science + Business

Media, New York, 2010, pp. 508-509.

iris-AperTO

University of Turin’s Institutional Research Information System and Open Access Institutional

Repository

(10)

[13] S. G. Lias, J. E. Bartmess, J. F. Liebman, J. L. Holmes, R. D. Levin, W. G. Mallard, Gas-phase ion and neutral

thermochemistry, J. Phys. Chem. Ref. Data, Suppl. 1 1988, 17, 1-861.

[14] NIST Chemistry WebBook, NIST Standard Reference Database Number 69. National Institute of Standards and Technology. Gaithersburg, MD, USA, June 2005, http://webbook.nist.gov.

[15] A. J. White, Chem. Soc. Rev. 1974, 3, 17-39.

[16] T.B. McMahon, T. Heinis, G. Nicol, J.K. Hovey, P. Kebarle, J. Am. Chem. Soc. 1988, 110, 7591-7598.

[17] V.V. Yakobson, D.G. Musaev, A.S. Zyubin, A.M. Mebel, O.P. Charkin, Koord Khim, Koord. Khim. 1989, 15, 1478-1488. [18] a) F. Grandinetti, J. Hrušàk, D. Schröder, S. Karrass, H. Schwarz, J. Am. Chem. Soc. 1992, 114, 2806-2810; b) F. Grandinetti, P. Cecchi, V. Vinciguerra, Chem. Phys. Lett. 1997, 281, 431-437; c) F. Grandinetti, V. Vinciguerra, J. Mol. Struct.

(Theochem) 2001, 574, 185-193; d) F. Grandinetti, V. Vinciguerra, Int. J. Mass Spectrom. 2002, 216, 285-299; e) M. Giordani,

F. Grandinetti, J. Fluorine Chem. 2009, 130, 557-561.

[19] K. Hiraoka, M. Nasu, S. Fujimaki, S. Yamabe, J. Phys. Chem. 1995, 99, 15822-15829.

[20] E.T. Denisov, O.M. Sarkisov and G.I. Likhtenshtein, Chemical Kinetics Fundamentals and New Developments, Elsevier Science B.V. Amsterdam, 2003, p. 381..

[21] Li-Qun Xia, M. Chang, in Handbook of semiconductor manufacturing technology 2th Ed., CRC Press, 2008, 13-1.

[22] C. Capellos, A. J. Swallow, J. Phys. Chem. 1969, 75, 1077-1083.

[23] Gaussian 09, Revision A.02, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K.

Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J.

Fox, Gaussian, Inc., Wallingford CT, 2009.

[24] MOLPRO, version 2010.1, a package of ab initio programs, H.-J. Werner, P. J. Knowles, F. R.Manby, M. Schütz, P. Celani, G. Knizia, T. Korona, R. Lindh, A. Mitrushenkov, G. Rauhut, T. B. Adler, R. D. Amos, A. Bernhardsson, A. Berning, D. L. Cooper, M. J. O. Deegan, A. J. Dobbyn, F. Eckert, E. Goll, C. Hampel, A. Hesselmann, G. Hetzer, T. Hrenar, G. Jansen, C. K¨oppl, Y. Liu, A. W. Lloyd, R. A. Mata, A. J. May, S. J. McNicholas, W. Meyer, M. E. Mura, A. Nicklass, P. Palmieri, K. Pflüger, R. Pitzer, M. Reiher, T. Shiozaki, H. Stoll, A. J. Stone, R. Tarroni, T. Thorsteinsson, M. Wang, and A. Wolf, , see http://www.molpro.net.

[25] H. B. Schlegel, J. Comput. Chem. 1982, 3, 214-218. [26] C. Møller, M. S. Plesset, Phys. Rev. 1934, 46, 618-622.

[27] B. Roos in Advances in Chemical Physics: Ab Initio Methods in Quantum Chemistry-II (Eds.: K.P. Lawley), Wiley, New York, 1987, pp.399-445. [31] H.B. Schlegel, J. Comput. Chem., 1982, 3, 214-218.

[28] W. J. Hehre, L. Radom, P. v. R. Schleyer, J. A. Pople, Ab initio Molecular Orbital Theory, Wiley, New York, 1986, p 80. [29] H. B. Schlegel, in Computational Theoretical Organic Chemistry, (Eds.: I. G. Csizmadia, R. Daudel), Reidel Publ Co, Holland, 1981, pp. 129-159.

[30] H. B. Schlegel, J. Chem. Phys. 1982, 77, 3676-3681.

[31] H. B. Schlegel, J. S. Binkley, J. A. Pople, J. Chem. Phys. 1984, 80, 1976-1981. [32] H. B. Schlegel, J. Comput. Chem. 1982, 3, 214-218.

[33] a) H. P. Hratchian, H. B. Schlegel, J. Chem. Phys. 2004, 120, 9918-9924; b) H. P. Hratchian, H. B. Schlegel, in Theory and

Applications of Computational Chemistry: The First 40 Years, (Eds.: C. E. Dykstra, G. Frenking, K. S. Kim, and G. Scuseria),

Elsevier, Amsterdam, 2005, pp. 195-249; c) H. P. Hratchian and H. B. Schlegel, J. Chem. Theory Comput. 2005, 1, 61-69. [34] D. A. Mc Quarry, Statistical Mechanics, Harper & Row, New York, 1976.

[35] K. Raghavachari, G. W. Trucks, J. A. Pople, M. Head-Gordon, Chem. Phys. Lett. 1989, 157, 479-483. [36] C. Hampel, K. Peterson, H. J. Werner, Chem. Phys. Lett. 1992, 190, 1-12.

[37] H.-J. Werner, P. J. Knowles, J. Chem. Phys. 1988, 89, 5803-5814. [38] P. J. Knowles, H.-J. Werner, Chem. Phys. Lett. 1988, 145, 514-522. [39] S. R. Langhoff, D. E. Davidson, Int. J. Quantum Chem. 1974, 8, 61-72. [40] T. J. Lee, P. R. Taylor, Int. J. Quantum Chem. 1989, 36, Suppl. S23, 199-207.

[41] L. A. Curtiss, K. Raghavachari, P. C. Redfern, V. Rassolov, J. A. Pople, J. Chem. Phys. 1998, 109, 7764-7776. [42] A. G. Baboul, L. A. Curtiss, P. C. Redfern, K. Raghavachari, J. Chem. Phys. 1999, 110, 7650 - 7657.

[43] R. Cooper, R. Mooring, Austr. J. Chem. 1968, 21, 2417-2425.

[44] I. Krajcar Bronic, Hoshasen (Ionizing Radiation) (Tokyo),1998, 24, 101-125.

Table of Contents

iris-AperTO

(11)

The reactions of alkyl radicals R∙ (R = CH3, C2H5, i-C3H7) at the

nitrogen atom of NF3 were studied

together with the competitive fluorine atom abstraction, by using X-ray irradiation and by computational methods at the CCSCD(T), MRCI and G3B3 levels of theory. The results indicate that the alkyl difluoroamines are produced by attack of alkyl radicals to the nitrogen atom of NF3

through a bimolecular homolytic substitution reaction mechanism.

Paola Benzi, Stefano Borocci, Chiara Demaria, Maria Giordani, Felice Grandinetti, Lorenza Operti, Roberto Rabezzana, Paola Antoniotti*

Page No. – Page No.

Bimolecular homolytic substitution at nitrogen: an experimental and theoretical study on the gas phase reactions of alkyl radicals with NF3

iris-AperTO

University of Turin’s Institutional Research Information System and Open Access Institutional

Repository

Riferimenti

Documenti correlati

Material and

Finally, we provide data on the production of female CHCs during the first days after emergence and, using hexane-washed CHC-free females or antennectomized males, conduct a

A fungal invasion is enhanced by hybridization and gene introgression: Ecological and evolutionary implications of genomic admixing.. Garbelotto

We have previously shown that whereas T-cells from normal individuals undergo accumulation of p53 and apoptosis when treated with the genotoxic agent Actinomycin D (ActD), those

Sono emerse delle riflessioni molto interessanti relativamente al tema della libertà ( più interessanti di quanto pensassi a priori). Ciascuno anche se con modalità

 Nella configurazione con barriere, invece, osserviamo che il camion si ribalta vicino la torre solo dopo una certa velocità del veicolo. Nel caso di veicolo

Fermo restando che il ricorso alle misure di sorveglianza elettronica a distanza non deve rappresentare o essere percepito come un alleggerimento della pena, il fatto

Stima terreni incolti 592.153,23 (337.107 proprietari) Superficie totale seminativi 6.938.830,68 ha (966.574 aziende) Stima superficie arboricoltura da legno 121.420 ha