0% found this document useful (0 votes)
23 views9 pages

Chem230Final Extended-Abstract (Edited)

The document discusses the synthesis and characterization of 3,5-diphenyl-2-phosphafuran (DPF) and its ability to undergo thermally reversible [4 + 2] cycloaddition reactions with various dienophiles. The synthesis involved the reaction of trans-chalcone with dibenzo-7-phosphanorbornadiene, yielding DPF in 43% yield, which was shown to effectively react with dimethyl acetylenedicarboxylate, norbornene, and ethylene. Analytical and computational studies confirmed DPF's potential as a diene in Diels–Alder reactions and its reversible binding with ethylene under mild heating.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOC, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
23 views9 pages

Chem230Final Extended-Abstract (Edited)

The document discusses the synthesis and characterization of 3,5-diphenyl-2-phosphafuran (DPF) and its ability to undergo thermally reversible [4 + 2] cycloaddition reactions with various dienophiles. The synthesis involved the reaction of trans-chalcone with dibenzo-7-phosphanorbornadiene, yielding DPF in 43% yield, which was shown to effectively react with dimethyl acetylenedicarboxylate, norbornene, and ethylene. Analytical and computational studies confirmed DPF's potential as a diene in Diels–Alder reactions and its reversible binding with ethylene under mild heating.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOC, PDF, TXT or read online on Scribd

3,5-Diphenyl-2-phosphafuran: Synthesis, Structure, and Thermally

Reversible [4 + 2] Cycloaddition Chemistry.

Martin-Louis Y. Riu1 and Christopher C. Cummins2†


1
Department of Chemistry
Institute of Technology, Cambridge, Massachusetts 02139;
orcid.org/0000-0002-0900-3545)
*Email: no email
2
Department of Chemistry,
Massachusetts Institute of Technology, Cambridge,
Massachusetts 02139; orcid.org/0000-0003-2568-3269
*Email: [email protected]

Presented by: Jhana Ruth V. Dano†



Department of Chemistry at Mindanao State University –
Iligan Institute of Technology, Iligan
*Email: [email protected]

Keywords: Diels–Alder; 3,5-diphenyl-2-phosphafuran (DPF); dienophile; diene; cycloaddition.

Researchers worked on the synthesis and characterization of a 2-phosphafuran and binds with
ethylene thermally under reversible cycloadditions. This was tested by treatment of trans-chalcone with
dibenzo-7-phosphanorbornadiene (EtOPA) (A = C14H10, anthracene), a source of ethoxyphosphinidene,
followed by elimination process of ethanol yields 3,5-diphenyl-2-phosphafuran (DPF) in 43% yield. Base on
the analytical and quantum computational results, the phosphadiene moiety of DPF is a potent diene in
the Diels–Alder reaction and as it reacts with dienophiles producing cycloadduct dimethyl
acetylenedicarboxylate (DPF·DMAD, 68%), norbornene (DPF·norbornene, 73%), and ethylene (DPF·C2H4,
80%) under ambient conditions. Also, a result showed DPF as a molecule that is able to reversibly bind
with ethylene through mild heating of DPF·C2H4 which significantly corresponds to retro-Diels–Alder
reaction.

1
Introduction determined the reaction pathways which
were consistent with the regio-selectivity
An increasing discovery through observed. in the reaction of styrene, and the
research and experiments of the heavier reaction was suggested to be essentially
main-group elements such as Al, Si, P has a concerted but highly asynchronous. [13,14] The
transition-metal like properties than the reaction of β-diketimine with AlR3 (R = Me,
lighter main-group congeners and many of Et, Bui) in hexanes yields title aluminum
the new compounds react with small complexes. The reaction could proceed with
molecules such as C2H4 or CO under mild β-hydride abstraction and reacts with
conditions and a potential application in ethylene to give 1,4-cycloaddn. [15, 16,17, 18]
catalysis. [1,2] Ethylene's cycloadditions to
unsaturated hydrocarbons have been An experiment on replacing carbon
develop and established literatures on by phosphorus in unsaturated molecules
classical organic chemistry. [3,4,5] One which can yield energetically accessible
examples are treatment of the distannynes frontier orbitals, due to the poor π overlap
with ethylene under ambient conditions between the carbon and phosphorus centers
[19]
affords the cycloadducts that were , this study aimed the targeted synthesis of
structurally and spectroscopically 2-phosphafuran and have tested its diene
characterized. Ethylene that incorporated potential to undergo low-barrier, thermally
with structures involves tin-carbon σ bonding reversible cycloadditions with some
and is fully reversible under ambient dienophile molecule. A thermally reversible
conditions; hydrocarbon solutions revert to cycloadditions between furan and maleimide
the distannynes compounds with ethylene moieties found from a number of furan
elimination under reduced pressure or upon groups that can repeatedly mend. [20] Some
standing at ~25°C. [3,5] Synthesis that studies shows also that phosphadienes reacts
involves cyclopropanation and epoxidation with activated dienophiles, such as
[21]
reactions were discovered to be rapid and maleimide. Related 1-
thermally reversible at relatively low phosphanorbornadienes studies have been
temperature. [6]Siliranes compounds are accessed by treating thermally generated
formed by reversible addition of ethylene to 2H-phospholes with dienophiles. [22−25] Taking
the corresponding stable phosphonium all the mentioned references, the study
silaylides which almost thermoneutral obtained two step preparation of 3,5-
addition with negative Gibb’s energy and diphenyl-2-phosphafuran or DPF as shown in
stable in low temperature and under ethylene Figure 1.
pressure (10 atm) in contrast to a usual high
exothermic cycloaddition of silylenes with
alkenes. [7,8,9,10] Another study was the heavier
group 13 element alkene analog, digallane
has been shown to react readily in [n + 2] (n
= 6, 4, 2 + 2) cycloaddition reactions with
norbornadiene and quadricyclane, 1,3,5,7-
cyclooctatetraene, 1,3-cyclopentadiene, and
1,3,5-cycloheptatriene to afford the heavier
element deltacyclane species and polycyclic Figure 1. General Structure of DPF.
compounds. [11] The synthesis process by
insertion of 1,1-diphenylethylene into the Mg- This DPF molecule has then reacted
Mg bond of two magnesium(I) dimers, with dienophiles and obtained results
yielding 1,2-dimagnesioethane products is supporting that its diene characteristic is a
also readily reversible at room temperature potent diene in Diels–Alder reaction. Details
and thus represent the first examples of of the overall experiment of the said
room-temperature reversible redox processes synthesis is stated in experimental method.
for s-block metal complexes. The said
product is highly activated magnesium alkyls Experimental Method
and show reactivity toward H2, CO, and
ethylene which supported by computational
Formation of 3,5-diphenyl-2-
studies. [12] Another cycloaddition reaction
phosphafuran (DPF) was achieved through
study a [4 + 2] cycloaddition reaction of
two steps reaction from heating a benzene
1,3,2,5-diazadiborinine with ethylene
solution of dibenzo-7-phosphanorbornadiene
afforded a bicyclo[2.2.2] under ambient
(EtOPA) and trans-chalcone. [27] The resulting
conditions, which the cyclization process was
mixture was then treated with sodium
reversible, and thus retro-[4 + 2]
bis(trimethylsilyl)amide (Na[N(SiMe3)2]) and
cycloaddition , with the release of ethylene.
trimethylsilyl chloride (Me3SiCl) to eliminate
Compound reacted regio-selectively and
HN(SiMe3)2, NaCl, and Me3SiOEt leaving the
stereo-selectively with styrene derivatives
DPF molecule. The produced DPF was reacted
and norbornene, and these processes were
with dienophile dimethyl
reversible too. Computational studies

2
acetylenedicarboxylate (DMAD), ethylene, mmol, 1.00 equiv; the of the) in 1,2-
and some olefins to accomplished Diels-Alder dimethoxyethane (20 mL) was added with
reaction. Overall process in achieving the sodium bis(trimethylsilyl)amide (0.323 g,
reaction is shown in Figure 1. 1.76 mmol, 1.00 equiv) portionwise in 1,2-
dimethoxyethane (20 mL). After vigorously
stirring for 15 min, the deep red solution was
frozen in the glovebox coldwell. To the
thawing solution was added trimethylsilyl
chloride (0.192 g, 1.76 mmol, 1.00 equiv)
dropwise and stirred as it warmed to 23°C in
1 hour until it became cloudy and bright
orange. All volatile materials were removed
in vacuo, and the resulting bright orange
solids were washed up in pentane (10 mL).
The Celite plug was washed with additional
pentane (10 mL) and all volatile materials
were removed from the combined filtrates in
Figure 2. Flow diagram in formation of Diels- vacuo, yielding a deep orange residue.
Alder adduct Crystallization from minimal pentane at −35
°C provided DPF as a pale orange crystalline
material (0.180 g, 1.51 mmol, 43%).
EtOPA
EtOPA trans-
trans-
chalcone As for the air stability of PDF, a 0.02
chalcone
M solution of DPF in benzene-d6 was
prepared and transferred to an NMR tube. To
DPF
DPF this tube was added a glass capillary
containing a 0.67 M solution of
triphenylphosphine in benzene-d6, and initial
NMR spectra were collected (Figures S.36
DPF.
DPF. DPF.C
DPF.C22H
H44 DPF.
DPF. and S.37). The cap of the tube was removed
DMAD
DMAD Norborene
Norborene
outside of the glovebox, and after 1 h the
solution became pale yellow.

The produced DPF was then treated


with different dienophile compound in which
Method used for 2-Ethoxy-3,5- reaction steps are of the following:
diphenyl-3-hydro-1,2-oxaphosphole crude
preparation, a solution of EtOPA (2.50 g, 9.84 1. Dimethyl acetylenedicarboxylate
mmol, 1.00 equiv) and trans-chalcone (2.05 solution (0.119 g, 0.840 mmol, 1.00 equiv)
g, 9.84 mmol, 1.00 equiv) in benzene (10 mL) was prepared in THF (2 mL), added dropwise
was prepared in a 25 mL Schlenk tube to a stirring solution of DPF (0.200 g, 0.840
containing a stir bar. The tube was removed mmol, 1.00 equiv) in THF (5 mL). After
from the glovebox and placed in an oil bath vigorous stirring for 20 min, all volatile
preheated to 80 °C. After 4 h, the flask was materials were removed in vacuo from the
removed from the oil bath and allowed to reaction mixture. A white heterogeneous
cool for 30 min, during which time flakes of mixture was obtained after adding pentane
anthracene crystallized. The sealed tube was (10 mL) resulting colorless residue and
returned to the glovebox, where all volatile stirring the solution for 20 min. The white
materials were removed in vacuo. To the precipitate was collected by vacuum filtration
residue was added Et2O (15 mL), and the using a 15 mL coarse sintered frit, and the
resulting slurry was filtered through a coarse solids were washed with minimal pentane
sintered frit (15 mL) containing a one-inch (ca. 2 × 5 mL) and dried to constant mass
plug of charcoal. The plug was washed with under reduced pressure. Crystallization from
Et2O (15 mL), and the combined filtrates minimal diethyl ether at −35 °C provided
were concentrated to 5 mL under reduced colorless crystals of DPF·DMAD (0.217 g,
pressure. The solution was then cooled to 0.571 mmol, 68%).
−35 °C in the glovebox freezer. After 24 h,
the supernatant was decanted away from the 2. For the preparation of 2,4-
white solids that had formed and all volatile Diphenyl-7-oxa-1-phosphabicyclo [2.2.1]
materials were removed from the hept-2-ene, DPF solution (100 mg, 0.42
supernatant under reduced pressure. mmol, 1 equiv) in THF (3 mL) was prepared in
the glovebox and transferred to a sealed 25
An attempt to separate 2 from the mL Schlenk tube containing a stir bar. The
mixture by crystallization was unsuccessful tube was removed from the glovebox,
thus proceeded to elimination of HOEt. degassed by three freeze–pump–thaw cycles,
Thawing solution of crude 2 (0.500 g, 1.76 and backfilled with ethylene (1.0 atm, 1.1

3
mmol, 0.031 g). After vigorously stirring for solution of triphenylphosphine. A 31P{1H}
12 h, the sealed tube was returned to the NMR spectrum of the solution was collected
glovebox, where all volatile materials were after 12 hours. Partial conversion (ca. 64%)
removed in vacuo. Pentane (10 mL) was of DPF to the corresponding cycloadduct, as a
added to the pale brown precipitate, and the mixture of diastereomers in a 3:8:25:62
resulting solution was filtered through a 15 molar ratio, was observed. Due to
mL coarse-sintered frit containing a one-inch overlapping signals in 1H NMR spectrum, we
plug of Celite. The plug was washed with are unable to assign the stereochemistry of
additional pentane (5 mL), and all volatile the cycloadducts.
materials were removed from the combined
filtrates under reduced pressure. X-ray Diffraction analysis for all the
Crystallization from minimal diethyl ether at product obtained, low-temperature diffraction
−35 °C provided colorless crystals of data collected on a Bruker-AXS X8 Kappa Duo
DPF·C2H4 (0.089 g, 0.33 mmol, 80%). diffractometer with IμS microsources,
coupled to a Photon 3 CPAD detector using
The thermal stability of DPF·C2H4 Mo Kα radiation (λ = 0.71073 Å) for the
was done by 0.03 M solution of DPF·C2H4 in structure of DPF·DMAD and a Smart APEX2
THF was prepared in the glovebox and CCD detector using Mo Kα radiation (λ =
transferred to a J Young tube containing a 0.71073 Å) for the structures of DPF and
flame-sealed glass capillary charged with a DPF·C2H4, performing ϕ- and ω-scans. The
0.67 M benzene-d6 solution of structures were solved by dual-space
triphenylphosphine. An initial 31P{1H} NMR methods using SHELXT [48] and refined
spectrum of the solution was collected. The against F2 on all data by full-matrix least-
tube was then placed in a preheated 65 °C oil squares with SHELXL-2017 [48] following
bath, and NMR spectra of the solution were established refinement strategies. [49,50] All
collected after 4 h and partial conversion of non-hydrogen atoms were refined
DPF·C2H4 to DPF was observed, as assessed anisotropically. All hydrogen atoms were
by 31P{1H} NMR spectroscopy. included into the model at geometrically
calculated positions and refined using a
Thermolysis of DPF·C2H4 (10 mg, riding model. The isotropic displacement
0.038 mmol, 1.0 equiv) in the presence of parameters of all hydrogen atoms were fixed
Norbornene solution (4 mg, 0.04 mmol, 1 to 1.2 times the U-value of the atoms they
equiv) in THF (0.7 mL) was prepared in the are linked to (1.5 times for methyl groups).
glovebox and transferred to a J Young tube
containing a flame-sealed glass capillary Single crystals of DPF were grown
charged with a 0.67 M benzene-d 6 solution of from −35 °C pentane, structure was solved in
triphenylphosphine an analyzed with the the orthorhombic space group Pnma with half
same NMR spectroscopy. The tube was then a molecule of DPF and no solvent molecules
placed in a preheated 75 °C oil bath, and in the asymmetric unit. The structure exhibits
31
P{1H} NMR spectra of the solution were whole molecule disorder about a
collected after 3 and 15 hours. crystallographic inversion center. Disorders
were refined with the help of the FLAT
3. Norbornene (0.040 g, 0.42 mmol, restraint in SHELXL and similarity restraints
1.0 equiv) was added to a solution of DPF on 1,2 and 1,3 distances. Similarity and rigid
(0.100 g, 0.420 mmol, 1.00 equiv) in THF (4 bond restraints for anisotropic displacement
mL) portionwise. After vigorous stirring for 5 parameters were applied to all non-hydrogen
h, all volatile materials were removed in atoms. As for DPF·DMAD were grown from
vacuo, resulting in a colorless residue. −35 °C diethyl ether. The structure was
Pentane (10 mL) was added to this residue, solved in the monoclinic space group P21/c
and the resulting solution was filtered with one molecule of DPF·DMAD and no
through a 15 mL coarse sintered frit solvent molecules in the asymmetric unit.
containing a one-inch plug of Celite. The plug The molecule exhibits no disorder. Similarity
was washed with additional pentane (5 mL), and rigid bond restraints for anisotropic
and all volatile materials were removed from displacement parameters were applied to all
the combined filtrates in vacuo. non-hydrogen atoms. For DPF·C2H4, crystal
Crystallization from minimal pentane at −35 was grown from −35 °C diethyl ether. The
°C provided colorless crystals of structure was solved in the monoclinic space
DPF·norbornene (0.102 g, 0.31 mmol, 73%). group P21/n with one molecule of DPF·C2H4
and no solvent molecules in the asymmetric
4. DPF was also reacted with 1- unit. The molecule exhibits no disorder.
Hexene, solution DPF (10 mg, 0.038 mmol, Similarity and rigid bond restraints for
1.0 equiv) and 1-hexene (32 mg, 0.38 mmol, anisotropic displacement parameters were
10 equiv) in benzene-d6 (0.7 mL) was applied to all non-hydrogen atoms.
prepared in the glovebox and transferred to
an NMR tube containing a flame-sealed glass Computational Studies
capillary charged with a 0.67 M benzene-d6

4
All calculations used by the
researchers were performed with the ORCA
4.0.1 quantum chemistry package from the
development team at the University of Bonn.
[51,52]
Initial geometries were constructed in
Avogadro. [53,54] Geometry optimizations were
performed at the ωB97X-D3/Def2-TZVPP level
of theory using keywords wB97X-D3 def2-
TZVPP def2/J TightSCF RIJCOSX Grid4 Figure 3: Minimum energy pathway for the
FinalGrid5 Opt for the geometries for EtOPA, reaction of EtOPA (1) with trans-chalcone to give
Trans-chalcone, EtOPA + trans-chalcone compound 2
(TS1), EtOPA + trans-chalcone (I1), I1 to
structure 2 (TS2), structure 2, and The graph suggests the cheletropic
Anthracene while the B3LYP-D3(BJ)/Def2- addition of trans-chalcone to EtOPA, resulting
TZVPP level using keywords B3LYP D3BJ in the formation of λ5-phosphorane I1 that
Def2-TZVPP TightSCF RIJCOSX Grid4 proceeded to TS1 with an activation barrier
FinalGrid5 Opt for the geometries of DPF, of +24.7 kcal/mol. It should be noted that
Ethylene, DPF + ethylene (TS), DPF·C2H4, attempts to locate a local minimum for a
and Parent 2-phosphafuran. For transition possible zwitterionic product of conjugate
states, the keyword Opt was replaced by addition consistently resulted in the
OptTS. Vibrational frequency calculations convergence to I1 and formation of it is
were carried out on the optimized geometries thermodynamically favorable and was not
using the keyword NumFreq. Intermediate found in NMR spectroscopy. It was also
and transition state geometries were found observed that this species is consistent with
to have zero and one imaginary frequency, the small barrier (+6.3 kcal/mol) to
respectively. Natural bond orbital (NBO) anthracene fragmentation supported with
calculations were performed using NBO6 [34] computations as concerted cheletropic
within the ORCA program. The natural extrusion pathway (TS2). [35−40]
resonance theory (NRT) keyword was
specified to generate natural resonance The attempt to separate oxo-3-
structures of the parent 2-phosphafuran phospholene from phospholene mixture by
molecule. Lower contributing resonance crystallization was unsuccessful. An
structures (all below 5%) were not included in elimination process of HOEt from oxo-3-
the main text. phospholene have carried out by first
deprotonation with sodium
Results and Discussion bis(trimethylsilyl)amide (Na[N(SiMe3)2]) and
the resulting in situ generated, deep red
A. Formation of DFP anion, was then treated with trimethylsilyl
Upon heating a solution of EtOPA (1) chloride (Me3SiCl), providing 3,5-diphenyl-2-
and trans-chalcone to 80 °C for 4 hours phosphafuran (DPF) upon elimination of
obtained a mixture diastereomers of oxo-3- HN(SiMe3)2, NaCl, and Me3SiOEt. Figure 4
phospholene shown in Figure 2 with 31P NMR shows the synthesis pathway of DPF.
δ 180.6 and δ 170.9 ppm; syn/anti ratio 1.5:1
which mixture and a small amount of EtOPA
and other byproducts are shown in Figure 2.

Figure 4: General Mechanism on Preparation of


DPF Proceeded in Two Steps from EtOPA (1)

Using NMR spectroscopy, DPF


Figure 2: Labeling scheme for oxo-3- displays a notably deshielded 31P{1H} NMR
phospholene mixture. resonance of δ 285.4 ppm in benzene.
Crystallization from minimal pentane at −35
From the quantum chemical °C obtained DPF, as pale orange crystals
calculations results, the energy landscape of obtaining 43% yield. Crystals of DPF suitable
formation of the major isomer of compound 2 for an X-ray diffraction study were grown
in Figure 4, the B3LYP-D3(BJ)/Def2-TZVPP from minimal diethyl ether at −35 44°C, and
level of DFT is shown on Figure 3. the molecular structure of DPF is shown in
Figure 5.

5
Results in DPF-dienophile mixture,
first with dimethyl acetylenedicarboxylate
(DMAD) at 23 °C results in the formation of
DPF·DMAD (31P NMR δ 95.3 ppm), and this
colorless crystal was isolated in 68% yield
grown from diethyl ether at −35 °C and
structurally characterized in a single-crystal
X-ray diffraction study shown in Figure 7
below.

Figure 5: A. Molecular structure of DPF shown with


50% probability thermal ellipsoids (H atoms are
omitted); B. Natural resonance theory (NRT) Figure 7: Molecular structure of cycloadducts
derived bond orders for DPF. C. Resonance weight DPF·DMAD (left) and DPF·C2H4 (right) shown with
calculated for the parent 2-phosphafuran using ellipsoids at the 50%
NRT. probability level. (H atom are omitted)

The experimental P–C bond length of DPF that was treated with
1.752(9) Å is between P–C single and double ethylene converting to DPF·C2H4 (31P NMR δ
bond lengths (1.86 and 1.69 Å, respectively). 94.5 ppm) was observed 12 h after exposure
[30]
The bond length is consistent with (1 atm) at 23 °C. A colorless crystals of
moderate π-delocalization of the DPF·C2H4 were grown from pentane at −35
phosphaalkene double bond. Using the °C in 80% yield. DPF·C2H4 was characterized
Cambridge Structural Database, no examples also by X-ray crystallography, [22−25] and its
of 2-phosphafuran compounds with which to molecular structure is also shown in Figure 7.
compare the structural data obtained for DPF This is opposite to the high temperatures and
as revealed in Figure 5. The major resonance pressures that is usually required for Diels–
structure yield 63.2% contributor to parent Alder addition of dienes to unactivated
molecule that resembles a diene, involving alkenes. [41-44]

localized C–C and C–P π-bonds, proposing for


Diels–Alder reaction with dienophiles. It was also found that treatment of
DPF with an excess of 1-hexene (10 equiv)
B. Diels- Alder reaction of DPF with led to partial conversion (ca. 64%) of DPF to
dienophiles the corresponding cycloadduct, as a mixture
of diastereomers in a 3:8:25:62 molar ratio,
With all the dienophile molecule after 12 h at 23 °C. However, no reaction was
reacted to the moiety phosphadiene of DPF, observed between DPF and unactivated
Figure 6 shows general Diels–Alder reaction internal olefins, such as cyclohexene and cis-
between DPF with dienophiles. 4-octene so as heating of these doesn’t show
any reversible reaction.
The significant formation of ethylene
and DPF was surprisingly detected in NMR
spectroscopy when DPF·C2H4 was heated to
65 °C in tetrahydrofuran. It was repeated at
75 °C in the presence of norbornene and a
clean formation of DPF·norbornene was
observed after 15 h. DPF·norbornene was
prepared independently to validate the
identity of the product, and yield 73% of it.
The reversible Diels–Alder reaction of
DPF with ethylene have investigated the
Figure 6: Diels–Alder Adducts of DPF reaction energy landscape using quantum
chemical calculations as shown in Figure 8.

6
Figure 9: General Mechanism of the Overall
Synthesis of Diels-Alder reaction in the experiment.
Figure 8: Calculated stationary points, transition
state (“TS”) and relative enthalpies (blue, kcal/mol)
and free energies
(pink, kcal/mol, 298.15 K) involved in the Diels– Acknowledgment
Alder addition of DPF to ethylene.
The researchers would like to thank
Calculations performed using DFT, at Michael B. Geeson, Wesley J. Transue, and Scott M.
the ωB97X-D3/Def2-TZVPP level of theory, Shepard for discussions consultations and
suggest that the cycloaddition is a concerted suggestions. The material is based on the research
process, proceeding with an activation supported by the National Science Foundation
under CHE-1955612.
barrier of +26.5 kcal/mol (TS). These
calculations also reveal that the reaction is
References
enthalpically favorable (ΔH = −18.1 kcal/mol)
but is downhill by only −4.8 kcal/mol,
1. Power, P. P. Main-group elements as
consistent with the experimental observation transition metals. Nature 2010, 463,
of retro cyclization at elevated temperatures. 171−177.
2. Martin, D.; Soleilhavoup, M.; Bertrand, G.
Summary and Conclusion Stable singlet carbenes as mimics for
transition metal centers. Chem. Sci. 2011,
2, 389−399.
The study introduces 3,5-diphenyl-2-
3. Peng, Y.; Ellis, B. D.; Wang, X.; Fettinger, J.
phosphafuran (DPF) as a potent diene in the CPower, P. Reversible Reactions of
Diels–Alder reaction. Showing all the relevant Ethylene with Distannynes Under Ambient
results mentioned, the contained Conditions. Science 2009, 325,
phosphadiene of the targeted synthesis of 1668−1670.
DPF displays deshielded 31P{1H} NMR 4. Hadlington, T. J.; Li, J.; Hermann, M.;
resonance of δ 285.4 ppm in benzene. The Davey, A.; Frenking, G.; Jones, C.
mechanism is supported with quantum Reactivity of Amido-Digermynes, LGeGeL
(L = Bulky Amide), toward Olefins and
calculations which proved the cheletropic
Related Molecules: Facile Reduction, C−H
addition of trans-chalcone to EtOPA resulting Activation, and Reversible Cycloaddition
the formation of λ5-phosphorane with an of Unsaturated Substrates.
activation barrier of +24.7 kcal/mol at its first Organometallics 2015, 34, 3175−3185.
transition sate (TS1) and the small barrier 5. Sugahara, T.; Guo, J.-D.; Sasamori, T.;
+6.3 kcal/mol of anthracene fragmentation Nagase, S.; Tokitoh, N. Reversible addition
from the ETOPA suggest the concerted of terminal alkenes to digermynes. Chem.
cheletropic extrusion pathway of second Commun. 2018, 54, 519−522.
6. Moerdyk, J. P.; Bielawski, C. W.
transition (TS2) forming the final product DPF Diamidocarbenes as versatile and
with 43% yield. The obtained DPF were reversible [2 + 1] cycloaddition reagents.
reacted with dienophiles, DMAD, ethylene Nat. Chem. 2012, 4, 275−280.
and norbornene forming Diels-Alder product: 7. Rodriguez, R.; Gau, D.; Kato, T.; Saffon-
DPF.DMAD (31P NMR δ 95.3 ppm), DPF.C2C4 Merceron, N.; De Cózar, A.; Cossío, F. P.;
(31P NMR δ 94.5 ppm) and DPF.norbornene Baceiredo, A. Reversible Binding of
that yield 68%, 80% and 73% respectively. Ethylene to Silylene − Phosphine
Complexes at Room Temperature. Angew.
Moreover, DPF was not only formed a
Chem., Int. Ed. 2011, 50, 10414−10416.
cycloadduct with ethylene, but also shows 8. Lips, F.; Fettinger, J. C.; Mansikkamäki, A.;
significant retro-Diels–Alder reaction where Tuononen, H. M.; Power, P. P. Reversible
the thermal decomposition of DPF.C2H4 Complexation of Ethylene by a Silylene
produced DPF and ethylene as detected in under Ambient Conditions. J. Am. Chem.
NMR. This is through 65 °C in tetrahydrofuran Soc. 2014, 136, 634−637.
and was repeated at 75 °C in the presence of 9. Lai, T. Y.; Gullett, K. L.; Chen, C.-Y.;
norbornene obtaining a clean NMR spectrum Fettinger, J. C.; Power, P. P. Reversible
Complexation of Alkynes by a Germylene.
of DPF.norbornene. The overall synthesis
Organometallics 2019, 38, 1421−1424.
result is summarized with mechanism shows 10. Sita, L. R.; Bickerstaff, R. D. Synthesis and
in Figure 9. crystal structure of the first

7
stannacyclopropene derivative. J. Am. 24. Le Goff, P.; Mathey, F.; Ricard, L. [4 + 2]
Chem. Soc. 1988, 110, 5208−5209. Cycloadditions between 2H-phospholes
11. Caputo, C. A.; Guo, J.-D.; Nagase, S.; and alkenes. Synthesis and properties of
Fettinger, J. C.; Power, P. P. Reversible and 1-phosphanorbornenes. J. Org. Chem.
Irreversible Higher-Order Cycloaddition 1989, 54, 4754−4758.
Reactions of Polyolefins with a Multiple- 25. Mathey, F. Transient 2H-Phospholes as
Bonded Heavier Group 13 Alkene Powerful Synthetic Intermediates in
Analogue: Contrasting the Behavior of Organophosphorus Chemistry. Acc. Chem.
Systems with π-π, π-π* and π-n+ Frontier Res. 2004, 37, 954−960.
Molecular Orbital Symmetry. J. Am. Chem. 26. Courtemanche, M.-A.; Transue, W. J.;
Soc. 2012, 134,7155−7164. Cummins, C. C. Phosphinidene Reactivity
12. Boutland, A. J.; Carroll, A.; Alvarez of a Transient Vanadium P≡N Complex. J.
Lamsfus, C.; Stasch, A.; Maron, L.; Jones, Am. Chem. Soc. 2016, 138,
C. Reversible Insertion of a C = C Bond 16220−16223.
into Magnesium(I) Dimers: Generation of 27. Transue, W. J.; Velian, A.; Nava, M.;
Highly Active 1,2-Dimagnesioethane García-Iriepa, C.; Temprado, M.; Cummins,
Compounds. J. Am. Chem. Soc. 2017, 139, C. C. Mechanism and Scope of
18190−18193. Phosphinidene Transfer from Dibenzo-7-
13. Wu, D.; Ganguly, R.; Li, Y.; Hoo, S. N.; phosphanorbornadiene Compounds. J. Am.
Hirao, H.; Kinjo, R. Reversible [4 + 2] Chem. Soc. 2017, 139, 10822−10831.
cycloaddition reaction of 1,3,2,5- 28. Duffy, M. P.; Lin, Y.; Ho, F.; Mathey, F.
diazadiborinine with ethylene. Chem. Sci. Synthesis and Chemistry of 2-
2015, 6, 7150−7155. Phosphafurans. Organometallics 2010,
14. Taylor, J. W.; McSkimming, A.; Guzman, C. 29, 5757−5758.
F.; Harman, W. H. N-Heterocyclic Carbene- Mack, A.; Bergsträßer, U.; Reiß, G. J.;
Stabilized Boranthrene as a Metal-Free Regitz, M. 3,5-Dimesityl-1,2,4-
Platform for the Activation of Small oxadiphosphole − Synthesis and
Molecules. J. Am. Chem. Soc. 2017, 139, Reactivity of a Novel Heterocycle. Eur. J.
11032−11035. Org. Chem. 1999, 1999, 587−595.
15. Radzewich, C. E.; Coles, M. P.; Jordan, R. F. 29. Mack, A.; Bergsträßer, U.; Reiß, G. J.;
Reversible Ethylene Cycloaddition Regitz, M. 3,5-Dimesityl-1,2,4-
Reactions of Cationic Aluminum β- oxadiphosphole − Synthesis and
Diketiminate Complexes. J. Am. Chem. Reactivity of a Novel Heterocycle. Eur. J.
Soc. 1998, 120, 9384−9385. Org. Chem. 1999, 1999, 587−595.
16. Bakewell, C.; White, A. J. P.; Crimmin, M. 30. Pyykkö, P.; Atsumi, M. Molecular Double-
R. Reactions of Fluoroalkenes with an Bond Covalent Radii for Elements
Aluminium(I) Complex. Angew. Chem., Int. Li−E112. Chem. - Eur. J. 2009, 15,
Ed. 2018, 57, 6638−6642. 12770−12779.
17. Schwamm, R. J.; Anker, M. D.; Lein, M.; 31. Glendening, E. D.; Weinhold, F. Natural
Coles, P. Reduction vs. Addition: The resonance theory: I. General formalism. J.
Reaction of an Aluminyl Anion with Comput. Chem. 1998, 19, 593−609.
1,3,5,7-Cyclooctatetraene. Angew. Chem., 32. Glendening, E. D.; Weinhold, F. Natural
Int. Ed. 2019, 58, 1489−1493. resonance theory: II. Natural bond order
18. Bakewell, C.; White, A. J. P.; Crimmin, M. and valency. J. Comput. Chem. 1998, 19,
R. Reversible alkene binding and allylic 610.
C−H activation with an aluminium(I) 33. Glendening, E. D.; Badenhoop, J. K.;
complex. Chem. Sci. 2019, 10, Weinhold, F. Natural resonance theory: III.
2452−2458. Chemical applications. J. Comput. Chem.
19. Kutzelnigg, W. Chemical Bonding in 1998,19, 628−646.
Higher Main Group Elements. Angew. 34. Glendening, E. D.; Badenhoop, J. K.; Reed,
Chem., Int. Ed. Engl. 1984, 23, 272−295. A. E.; Carpenter, J. E.; Bohmann, J. A.;
20. Chen, X.; Dam, M. A.; Ono, K.; Mal, A.; Morales, C. M.; Landis, C. R.; Weinhold, F.
Shen, H.; Nutt, S. R. Sheran, K.; Wudl, F. A NBO 6.0; Theoretical Chemistry Institute,
Thermally Re-mendable Cross-Linked University of Wisconsin: Madison, WI,
Polymeric Material. Science 2002, 295, 2013.
1698−1702. 35. Mesch, K. A.; Quin, D. Syn to Anti
21. Liu, Y.-L.; Chuo, T.-W. Self-healing Isomerization and Degration of
polymers based on thermally reversible Phosphines of the 7-Phosphanorbornene
Diels−Alder chemistry. Polym. Chem. System by Action of Methanol.
2013, 4, 2194−2205. Tetrahedron Lett. 1980, 21, 4791−4794.
22. Mathey, F.; Mercier, F.; Charrier, C.; 36. Newkome, G. R.; Hager, D. C. A new
Fischer, J.; Mitschler, A. Dicoordinated 2H- contractive coupling procedure.
phospholes as transient intermediates in Convenient phosphorus expulsion
the reactions of tervalent phospholes at reaction. J. Am. Chem. Soc. 1978, 100,
high temperature. One-step syntheses of 5567−5568.
1-phosphanorbornadienes and 37. Uchida, Y.; Onoue, K.; Tada, N.; Nagao, F.;
phosphorins from phospholes. J. Am. Oae, S. Ligand coupling reaction on the
Chem. Soc. 1981, 103, 4595−4597. phosphorus atom. Tetrahedron Lett.
23. Charrier, C.; Bonnard, H.; De Lauzon, G.; 1989, 30, 567−570.
Mathey, F. Proton [1,5] shifts in P- 38. Uchida, Y.; Kozawa, H. Formation of 2,2′-
unsubstituted 1H-phospholes. Synthesis bipyridyl by ligand coupling on the
and chemistry of 2H-phosphole dimers. J. phosphorus atom. Tetrahedron Lett.
Am. Chem. Soc. 1983, 105, 6871−6877. 1989, 30, 6365−6368.

8
39. Hilton, M. C.; Zhang, X.; Boyle, B. T.;
Alegre-Requena, J. V.; Paton, R. S.;
McNally, A. Heterobiaryl synthesis by
contractive C−C coupling via P(V)
intermediates. Science 2018, 362,
799−804.
40. Szkop, K. M.; Geeson, M. B.; Stephan, D.
W.; Cummins, C. C. Synthesis of
acyl(chloro)phosphines enabled by
phosphinidene transfer. Chem. Sci. 2019,
10, 3627−3631.
41. Sauer, J.; Sustmann, R. Mechanistic
Aspects of Diels-Alder Reactions: A Critical
Survey. Angew. Chem., Int. Ed. Engl.
1980, 19, 779−807.
42. Tobia, D.; Harrison, R.; Phillips, B.; White,
T. L.; DiMare, M.; Rickborn, B. Unusual
stability of N-methylmaleimide
cycloadducts: characterization of
isobenzofuran retro-Diels-Alder reactions.
J. Org. Chem. 1993, 58, 6701−6706.
43. M. E.; Nicolaou, K. C. The Oxide Anion
Accelerated Retro-Diels-Alder Reaction.
Chem. - Eur. J. 1997, 3, 187−192.
44. Kotha, S.; Banerjee, S. Recent
developments in the retro-Diels-Alder
reaction. RSC Adv. 2013, 3, 7642−7666.
45. Pangborn, A. B.; Giardello, M. A.; Grubbs,
R. H.; Rosen, R. K.; Timmers, F. J. Safe and
Convenient Procedure for Solvent
Purification. Organometallics 1996, 15,
1518−1520.
46. Williams, D. B. G.; Lawton, M. Drying of
Organic Solvents: Quantitative Evaluation
of the Efficiency of Several Desiccants. J.
Org. Chem. 2010, 75, 8351−8354. Fulmer,
G. R.; Miller, A. J. M.; Sherden, N. H.;
Gottlieb, H. E.;
47. Nudelman, A.; Stoltz, B. M.; Bercaw, J. E.;
Goldberg, K. I. NMR Chemical Shifts of
Trace Impurities: Common Laboratory
Solvents, Organics, and Gases in
Deuterated Solvents Relevant to th
Organometallic Chemist. Organometallics
2010, 29, 2176−2179.
48. Sheldrick, G. M. SHELXT − Integrated
space-group and crystal-structure
determination. Acta Crystallogr., Sect. A:
Found. Adv. 2015, 71, 3−8.
49. Müller, P.; Herbst-Irmer, R.; Spek, A. L.;
Schneider, T. R.; Sawaya, M. R. Crystal
Structure Refinement: A
Crystallographer’s Guide to SHELXL; IUCr
Texts on Crystallography; Oxford
University Press: Oxford, 2006.
50. Müller, P. Practical suggestions for better
crystal structures. Crystallogr. Rev. 2009,
15, 57−83.
51. Neese, F. The ORCA program system.
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
2012, 2, 73−78.
52. Neese, F. Software update: the ORCA
program system, version 4.0. Wiley
Interdiscip. Rev.: Comput. Mol. Sci. 2018,
8, e1327.
53. Avogadro: an open-source molecular
builder and visualization tool. Version
1.2.0. http://avogadro.cc/.
54. Hanwell, M. D.; Curtis, D. E.; Lonie, D. C.;
Vandermeersch, T.; Zurek, E.; Hutchison,
G. R. Avogadro: an advanced semantic
chemical editor, visualization, and
analysis platform. J. Cheminf. 2012, 4, 17.

You might also like