20 research outputs found
Carbenic Nitrile Imines: Properties and Reactivity
Structures and properties of nitrile
imines were investigated computationally
at B3LYP and CCSD(T) levels. Whereas NBO analysis at the B3LYP DFT
level invariably predicts a propargylic electronic structure, CCSD(T)
calculations permit a clear distinction between propargylic, allenic,
and carbenic structures. Nitrile imines with strong IR absorptions
above ca. 2150 cm<sup>–1</sup> have propargylic structures
with a CN triple bond (RCNNSiMe<sub>3</sub> and R<sub>2</sub>BCNNBR<sub>2</sub>), and those with IR absorptions below ca. 2150 cm<sup>–1</sup> are allenic (HCNNH, PhCNNH, and HCNNPh). Nitrile imines lacking
significant cumulenic IR absorptions at 1900–2200 cm<sup>–1</sup> are carbenic (R–(<i>C</i>:)–NN–R′).
Electronegative but lone pair-donating groups NR<sub>2</sub>, OR,
and F stabilize the carbenic form of nitrile imines in the same way
they stabilize “normal” singlet carbenes, including <i>N</i>-heterocyclic carbenes. NBO analyses at the CCSD(T) level
confirm the classification into propargylic, allenic, and carbenic
reactivity types. Carbenic nitrile imines are predicted to form azoketenes <b>21</b> with CO, to form [2+2] and [2+4] cycloadducts and borane
adducts, and to cyclize to 1<i>H-</i>diazirenes of the type <b>24</b> in mildly exothermic reactions with activation energies
in the range 29–38 kcal/mol. Such reactions will be readily
accessible photochemically and thermally, e.g., under the conditions
of matrix photolysis and flash vacuum thermolysis
Rearrangements of Acyl, Thioacyl, and Imidoyl (Thio)cyanates to Iso(thio)cyanates, Acyl Iso(thio)cyanates to (Thio)acyl Isocyanates, and Imidoyl Iso(thio)cyanates to (Thio)acyl Carbodiimides, RCX-YCN ⇌ RCX-NCY ⇌ RCY-NCX ⇌ RCY-XCN (X and Y = O, S, NR′)
Two types of rearrangements have been investigated computationally
at the B3LYP/6-311+G(d,p) level. The activation barriers for rearrangement
of acyl thiocyanates RCO–SCN to the corresponding isothiocyanates
RCO–NCS are 30–31 kcal/mol in agreement with the observation
that the thiocyanates are in some cases isolable albeit very sensitive
compounds. Alkoxycarbonyl-, (alkylthio)carbonyl- and carbamoyl thiocyanates
are isolable and have higher calculated barriers (ca. 40 kcal/mol)
toward rearrangement to isothiocyanates, whereas all thioacyl thiocyanate
derivatives are rather unstable compounds with barriers in the range
20–30 kcal/mol for rearrangement to the isothiocyanates. Acyl-,
alkoxycarbonyl-, and carbamoyl cyanates R–CO–OCN are
predicted to be in some cases isolable compounds with barriers up
to ca. 40 kcal/mol for rearrangement to the isocyanates RCO–NCO.
All of the rearrangements of this type involve the HOMO of a nearly
linear (thio)cyanate anion and the LUMO of the acyl cation, in particular
the acyl CX π* orbital. The second type of rearrangement
involves 1,3-shifts of the groups R attached to the (thio)acyl groups,
that is, acyl isothiocyanate–thioacyl isocyanate and imidoyl
isothiocyanate–thioacyl carbodiimide rearrangements. These
reactions involve four-membered cyclic, zwitterionic transition states
facilitated by lone pair–LUMO interactions between the migrating
R group and the neighboring iso(thio)cyanate function. Combination
of the two rearrangements leads to the general reaction scheme RCX–YCN
⇌ RCX–NCY ⇌ RCY–NCX ⇌ RCY–XCN
(X and Y = O, S, NR′)
Formation of (Aza)fulvenallene, Cyanocyclopentadiene, and (Aza)fluorenes in the Thermal Rearrangements of Indazoles, Azaindazoles, and Homoquinolinic Anhydride
Flash vacuum pyrolysis (FVP) of pyrazoles and indazoles
constitutes
a valuable route to carbenes and nitrenes. In this study, we employed
M062X and CCSD(T) calculations to provide a mechanistic rationale
for the formation of fulvenallene and fluorenes from indazoles and
the corresponding formation of azafulvenallene 15, cyanocyclopentadiene 19, and azafluorenes, e.g. 45, from azaindazoles,
e.g. 12, and from homoquinolinic anhydride. The results
reveal the importance of initial tautomerization in the pyrazole moiety
of 7-azaindazole 12, which drives the mechanism toward
2-diazo-3-methylene-2,3-dihydropyridine 29 and hence
3-methylene-2,3-dihydropyridin-2-ylidene 26, followed
by Wolff-type ring contraction to 1-azafulvenallene 15. This path has a calculated activation energy ∼10 kcal/mol
lower than that for an alternate route involving ring opening to 3-diazomethylpyridine,
dediazotization, and rearrangement of 3-pyridylcarbene to azacycloheptatetraene
and phenylnitrene 24. FVP of 2,5-diphenyltetrazoles and
phenyl(pyridyl)tetrazoles leads to nitrile imines, which cyclize to
3-phenylindazoles and -azaindazoles. Nitrogen elimination from these
(aza) indazoles results in the formation of (aza) fluorenes, for which
two alternate mechanisms are described: route A by rearrangement of
(aza) indazoles to diazo(aza)cyclohexadienes and (aza)cyclohexadienylidenes
and route B by rearrangement to diaryldiazomethanes and diarylcarbenes.
Both paths are energetically feasible, but path A is preferred and
corresponds to the azafluorenes obtained experimentally
Mechanistic Diversity in Thermal Fragmentation Reactions: A Computational Exploration of CO and CO<sub>2</sub> Extrusions from Five-Membered Rings
The
mechanisms of a variety of thermal pericyclic fragmentation reactions
of five-membered heterocyclic rings are subjected to scrutiny at a
density functional level by computation of transition state free energy
barriers and intrinsic reaction coordinates (IRCs). The preferred
computed products generally match those observed in flash vacuum thermolysis
experiments. For certain reactions, which also have the highest reaction
temperatures and computed barriers, a degree of multireference character
to the wave function manifests in an overestimation of the DFT-computed
barrier, with a more reasonable barrier obtained by a CASSCF single
point energy calculation. Many of the IRCs exhibit “hidden
intermediates” along the reaction pathway, but conversely reactions
that could be considered to involve the formation of an intermediate
nitrene prior to alkyl or aryl migration show no evidence of such
an intermediate. Such exploration of the diversity of behavior in
a class of compounds using computational methods with interactive
presentation of the results within the body of a journal article is
suggested as being almost a <i>sine qua non</i> for laboratory-based
research on reactive intermediates
Laser-Induced Carbene–Carbene Rearrangement in Solution: The Diphenylcarbene–Fluorene Rearrangement
Diphenylcarbene
(DPC) generated by high-intensity laser photolysis
of diphenyldiazomethane rearranges to fluorene (FL) by two distinct
mechanisms as revealed by methyl-group labeling. Thus, excimer laser
irradiation of <i>p</i>,<i>p</i>′-dimethyldiphenyldiazomethane
generates 3,6-dimethylfluorene (3,6-DMF) and 2,7-dimethylfluorene
(2,7-DMF), which were identified by fluorescence measurements as well
as GC-MS and comparison with authentic materials. 3,6-DMF corresponds
to direct bond formation between <i>ortho</i> positions
in DPC, referred to as <i>ortho</i>,<i>ortho</i>′ coupling. 2,7-DMF corresponds to a carbene–carbene
rearrangement, whereby DPC undergoes ring expansion to phenylcycloheptatetraene
(PhCHT) followed by ring contraction to <i>o</i>-biphenylylcarbene
(<i>o</i>-BPC), which then cyclizes to FL. The carbene–carbene
rearrangement dominates over the <i>ortho</i>,<i>ortho</i>′ coupling under all conditions employed. The <i>ortho</i>,<i>ortho</i>′ coupling must take place in a higher
excited state (most likely S<sub>2</sub> or T<sub>1</sub>) of DPC,
because it is not observed at all under thermolysis conditions, where
only S<sub>1</sub> and T<sub>0</sub> are populated. The carbene–carbene
rearrangement may take place either in a hot S<sub>1</sub> state or
more likely in a higher excited state (S<sub>2</sub> or T<sub>1</sub>)
3‑Pyridazinylnitrenes and 2‑Pyrimidinylnitrenes
Mild
flash vacuum thermolysis of tetrazolo[1,5-<i>b</i>]pyridazines <b>8T</b> generates small amounts of 3-azidopyridazines <b>8A</b> (<b>8aA</b>, IR 2145, 2118 cm<sup>–1</sup>; <b>8bA</b>, 2142 cm<sup>–1</sup>). Photolysis of the
tetrazoles/azides <b>8T/8A</b> in Ar matrix generates 3-pyridazinylnitrenes <b>9</b>, detected by ESR spectroscopy (<b>9a</b>: <i>D</i>/<i>hc</i> = 1.006; <i>E</i>/<i>hc</i> = 0.003 cm<sup>–1</sup>). Cyanovinylcarbenes <b>11</b>, derived from 4-diazobut-2-enenitriles <b>10</b>,
are also detected by ESR spectroscopy (<b>11a</b>: <i>D</i>/<i>hc</i> = 0.362; <i>E</i>/<i>hc</i> = 0.021 cm<sup>–1</sup>). Carbenes <b>11</b> rearrange
to cyanoallenes <b>12</b> and 3-cyanocyclopropenes <b>13</b>. Triazacycloheptatetraenes <b>20</b> were not observed in the photolyses of <b>8</b>. Photolysis of tetrazolo[1,5-<i>a</i>]pyrimidines/2-azidopyridmidines <b>18T/18A</b> in Ar matrices at 254 nm yields 2-pyrimidinylnitrenes <b>19</b>, observable by ESR, UV, and IR spectroscopy (<b>19a</b>: ESR: <i>D</i>/<i>hc</i> = 1.217; <i>E</i>/<i>hc</i> = 0.0052 cm<sup>–1</sup>). Excellent
agreement with the calculated IR spectrum identifies the 1,2,4-triazacyclohepta-1,2,4,6-tetraenes <b>20</b> (<b>20a</b>, 1969 cm<sup>–1</sup>; <b>20b</b>, 1979 cm<sup>–1</sup>). Compounds <b>20</b> undergo
photochemical ring-opening to 1-isocyano-3-diazopropenes <b>23</b>. Further irradiation also causes Type II ring-opening of pyrimidinylnitrenes <b>19</b> to 2-(cyanimino)vinylnitrenes <b>21</b> (<b>21a</b>: <i>D</i>/<i>hc</i> = 0.875; <i>E</i>/<i>hc</i> = 0.00 cm<sup>–1</sup>), isomerization
to cyaniminoketenimine <b>25</b> (2044 cm<sup>–1</sup>), and cyclization to 1-cyanopyrazoles <b>22</b>. The reaction
mechanisms are discussed and supported by DFT calculations on key
intermediates and pathways. There is no evidence for the interconversion
of 3-pyridazinylnitrenes <b>9</b> and 2-pyrimidinylnitrenes <b>19</b>
Nitrene-Carbene-Carbene Rearrangement. Photolysis and Thermolysis of Tetrazolo[5,1‑<i>a</i>]phthalazine with Formation of 1‑Phthalazinylnitrene, <i>o-</i>Cyanophenylcarbene, and Phenylcyanocarbene
1-Azidophthalazine <b>9A</b> is generated in trace amount
by mild FVT of tetrazolo[5,1<i>-a</i>]phthalazine <b>9T</b> and is observable by its absorption at 2121 cm<sup>–1</sup> in the Ar matrix IR spectrum. Ar matrix photolysis of <b>9T/9A</b> at 254 nm causes ring opening to generate two conformers of (<i>o-</i>cyanophenyl)diazomethane <b>11</b> (2079 and 2075
cm<sup>–1</sup>), followed by (<i>o</i>-cyanophenyl)carbene <sup>3</sup><b>12</b>, cyanocycloheptatetraene <b>13</b>,
and finally cyano(phenyl)carbene <sup>3</sup><b>14</b> as evaluated
by IR spectroscopy. The two carbenes <sup>3</sup><b>12</b> and <sup>3</sup><b>14</b> were observed by ESR spectroscopy (<i>D</i>|<i>hc</i> = 0.5078, <i>E</i>|<i>hc</i> = 0.0236 and <i>D</i>|<i>hc</i> =
0.6488, <i>E</i>|<i>hc</i> = 0.0195 cm<sup>–1</sup>, respectively). The rearrangement of <b>12</b> ⇄ <b>13</b> ⇄ <b>14</b> constitutes a carbene–carbene
rearrangement. 1-Phthalazinylnitrene <sup>3</sup><b>10</b> is
observed by means of its UV–vis spectrum in Ar matrix following
FVT of <b>9</b> above 550 °C. Rearrangement to cyanophenylcarbenes
also takes place on FVT of <b>9</b> as evidenced by observation
of the products of ring contraction, viz., fulvenallenes and ethynylcyclopentadienes <b>16</b>–<b>18</b>. Thus the overall rearrangement <b>10</b> → <b>11 → 12</b> ⇄ <b>13</b> ⇄ <b>14</b> can be formulated
C<sub>15</sub>H<sub>10</sub> and C<sub>15</sub>H<sub>12</sub> Thermal Chemistry: Phenanthrylcarbene Isomers and Phenylindenes by Falling Solid Flash Vacuum Pyrolysis of Tetrazoles
2-Phenyl-5-(phenylethynyl)tetrazole <b>44</b> provides a
new entry to the C<sub>15</sub>H<sub>10</sub> energy surface. Flash
vacuum pyrolysis of <b>44</b> using the falling solid flash
vacuum pyrolysis (FS-FVP) method afforded cyclopenta[<i>def</i>]phenanthrene <b>31</b> and cyclopenta[<i>jk</i>]fluorene <b>52</b> as the principal products. The products are explained in
terms of the formation of <i>N-</i>phenyl-<i>C</i>-phenylethynylnitrile imine/(phenylazo)(phenylethynyl)carbene <b>45</b> and its cyclization to 3-(phenylethynyl)-3<i>H-</i>indazole <b>46b</b>. Pyrolytic loss of N<sub>2</sub> from <b>46b</b> generates C<sub>15</sub>H<sub>10</sub> intermediate <b>48</b>. Cyclization of <b>48</b> to a dibenzocycloheptatetraene
derivative and further rearrangements with analogies in the chemistry
of phenylcarbene and the naphthylcarbenes leads to the final products.
Similar pyrolysis of 2-phenyl-5-styryltetrazole <b>43</b> afforded
3-styrylindazole <b>58</b>, which on further pyrolysis eliminated
N<sub>2</sub> to generate 3- and 2-phenylindenes <b>61</b> and <b>62</b> via C<sub>15</sub>H<sub>12</sub> intermediates
Nitrile Imines: Matrix Isolation, IR Spectra, Structures, and Rearrangement to Carbodiimides
The structures and reactivities of nitrile imines are
subjects of continuing debate. Several nitrile imines were generated
photochemically or thermally and investigated by IR spectroscopy in
Ar matrices at cryogenic temperatures (Ph-CNN-H <b>6</b>, Ph-CNN-CH<sub>3</sub> <b>17</b>, Ph-CNN-SiMe<sub>3</sub> <b>23</b>,
Ph-CNN-Ph <b>29</b>, Ph<sub>3</sub>C-CNN-CPh<sub>3</sub> <b>34</b>, and the boryl-CNN-boryl derivative <b>39</b>). The
effect of substituents on the structures and IR absorptions of nitrile
imines was investigated computationally at the B3LYP/6-31G* level.
IR spectra were analyzed in terms of calculated anharmonic vibrational
spectra and were generally in very good agreement with the calculated
spectra. Infrared spectra were found to reflect the structures of
nitrile imines accurately. Nitrile imines with IR absorptions above
2200 cm<sup>–1</sup> have essentially propargylic structures,
possessing a CN triple bond (typically PhCNNSiMe<sub>3</sub> <b>23</b>, PhCNNPh <b>29</b>, and boryl-CNN-boryl <b>39</b>). Nitrile imines with IR absorptions below ca. 2200 cm<sup>–1</sup> are more likely to be allenic (e.g., HCNNH <b>1</b>, PhCNNH <b>6</b>, HCNNPh <b>43</b>, PhCNNCH<sub>3</sub> <b>17</b>, and Ph<sub>3</sub>C-CNN-CPh<sub>3</sub> <b>34</b>). All nitrile
imines isomerize to the corresponding carbodiimides both thermally
and photochemically. Monosubstituted carbodiimides isomerize thermally
to the corresponding cyanamides (e.g., Ph-NCN-H <b>5</b> → Ph-NH-CN <b>8</b>), which are therefore the
thermal end products for nitrile imines of the types RCNNH and HCNNR.
This tautomerization is reversible under flash vacuum thermolysis
conditions
Ketene–Ketene Interconversion. 6‑Carbonylcyclohexa-2,4-dienone–Hepta-1,2,4,6-tetraene-1,7-dione–6-Oxocyclohexa-2,4-dienylidene and Wolff Rearrangement to Fulven-6-one
6-Carbonylcyclohexa-2,4-dienone (<b>1</b>) has been generated
by flash vacuum thermolysis (FVT) with Ar-matrix isolation of methyl
salicylate (<b>7</b>), 2-phenylbenzo-1,3-dioxan-4-one (<b>8</b>), phthalic peranhydride (<b>9</b>), and benzofuran-2,3-dione
(<b>11</b>) and also by matrix photolysis of <b>9</b>, <b>11</b>, and 2-diazocyclohepta-4,6-dien-1,3-dione (<b>12</b>). In each case, FVT above 600 °C results in decarbonylation
of <b>1</b> and Wolff rearrangement to fulven-6-one (<b>13</b>) either concertedly or via open-shell singlet 6-oxocyclohexa-2,4-dienylidene
(<b>18</b>). Ketenes <b>1</b> and <b>13</b> were
characterized by IR spectroscopy. Photolysis of matrix-isolated <b>1</b> at 254 nm also results in the slow formation of <b>13</b>. The sequential formation of ketenes <b>1</b> and <b>13</b> from <b>7</b> has also been monitored by FVT-mass spectrometry,
and <b>13</b> has been trapped with MeOH to afford methyl 1,3-cyclopentadiene-1-
and -2-carboxylates <b>15</b> and <b>16</b>. FVT of methyl
salicylate-1-<sup>13</sup>C <b>7a</b> revealed a deep-seated
rearrangement of the <sup>13</sup>C-labeled <b>1a</b> to hepta-1,2,4,6-tetraen-1,7-dione
(<b>17a</b>) by means of electrocyclic ring opening followed
by a facile 1,5-H shift and recyclization prior to CO-elimination
and ring contraction to <sup>13</sup>C-labeled <b>13</b>. The
rearrangement mechanism is supported by M06-2X/6-311++G(d,p) calculations,
which predict feasible barriers for the FVT rearrangements and confirm
the observed labeling pattern in the isolated methyl salicylate <b>7a/7b</b> and methyl cyclopentadienecarboxylates <b>20</b> and <b>21</b> resulting from trapping of <b>13</b> with
MeOH